09.02.2013 Views

click here to download ISOPOL XVII Book of - ISOPOL - International ...

click here to download ISOPOL XVII Book of - ISOPOL - International ...

click here to download ISOPOL XVII Book of - ISOPOL - International ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>International</strong> Symposium on Problems <strong>of</strong> Listeriosis


Ficha Técnica<br />

LIVRO DE ACTAS DO CONGRESSO: <strong>ISOPOL</strong> <strong>XVII</strong><br />

Intrenational Symposium on Problems <strong>of</strong> Listeriosis<br />

Edi<strong>to</strong>r: Universidade Católica Portuguesa – Escola Superior de Biotecnologia<br />

Coordenação e Revisão: Paula Teixeira<br />

Design e Composição Gráfica: Kai Sprecher / Lynn Salt<br />

Impressão: Orgal Impressores<br />

Depósi<strong>to</strong> Legal:<br />

Tiragem: 500 exemplares


WELCOME TO <strong>ISOPOL</strong> <strong>XVII</strong><br />

On behalf <strong>of</strong> the Organising Committee, I am pleased <strong>to</strong> welcome you <strong>to</strong><br />

<strong>ISOPOL</strong> <strong>XVII</strong> (<strong>International</strong> Symposium On Problems Of Listeriosis) which<br />

this year is organized in Por<strong>to</strong> by the Universidade Católica Portuguesa –<br />

Escola Superior de Biotecnologia.<br />

The <strong>ISOPOL</strong> meetings provide a unique opportunity for interdisciplinary<br />

discussions concerning various aspects <strong>of</strong> listeriosis including, but not being<br />

limited <strong>to</strong>, food safety and clinical aspects. Since the first edition <strong>of</strong> this<br />

symposium in Giessen, Germany (1957), much has been discovered and<br />

many more questions have been posed. For this reason the underlying motives<br />

for bringing <strong>to</strong>gether the evolving <strong>ISOPOL</strong> community are still very<br />

strong in 2010.<br />

More than 350 delegates from the clinical, veterinary, and public health<br />

areas as well as the food industry, will be present, representing ca. 45 countries.<br />

This diversity will certainly provide for a very stimulating and plural<br />

atmosp<strong>here</strong> which is sure <strong>to</strong> promote rich discussions and exchanges <strong>of</strong><br />

ideas.<br />

On this occasion, we would like <strong>to</strong> thank <strong>to</strong> the members <strong>of</strong> the Scientific<br />

Committee for their time and effort in maintaining the high scientific level<br />

<strong>of</strong> <strong>ISOPOL</strong> <strong>XVII</strong>.<br />

We are also thankful <strong>to</strong> our sponsors, which, in many different ways, have<br />

greatly contributed <strong>to</strong> the success <strong>of</strong> the organization <strong>of</strong> this event.<br />

For the first time ever, <strong>ISOPOL</strong> will take place in Portugal, namely in Por<strong>to</strong><br />

– the country’s second city and the capital <strong>of</strong> the north.<br />

It is easy <strong>to</strong> see why Por<strong>to</strong> was designated a World Heritage Site by UN-<br />

ESCO in 1996. It is a living city, full <strong>of</strong> monuments and signs <strong>of</strong> its rich his<strong>to</strong>ry.<br />

It is best met in person; take the opportunity <strong>to</strong> stay awhile, <strong>to</strong> meet<br />

the people, <strong>to</strong> enjoy the charms <strong>of</strong> this city <strong>of</strong> contrasts – you may find the<br />

need <strong>to</strong> return again <strong>to</strong> enjoy it all!<br />

Paula Cristina Maia Teixeira


TABLE OF CONTENTS<br />

KEY NOTE SPEECH<br />

Listeria monocy<strong>to</strong>genes: a multifaceted model 21<br />

Cossart, P.<br />

AREA //A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

PLENARY LECTURES<br />

A / PL/ 01 The pangenome <strong>of</strong> Listera spp. 22<br />

Chakraborty, T.<br />

A / PL/ 02 Ecology <strong>of</strong> L. monocy<strong>to</strong>genes and Listeria spp. in natural 22<br />

and food associated environments<br />

Wiedmann, M.<br />

ORAL PRESENTATIONS<br />

A / O/ 01 The cold shock associated proteins (Csps) promote <strong>to</strong>lerance 33<br />

<strong>of</strong> different environmental stresses and host cell invasion<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Tasara, T., Klumpp, J., Loessner, M. J. and Stephan, R.<br />

A / O/ 02 Different levels <strong>of</strong> flagellin detected during growth <strong>of</strong> 33<br />

Listeria monocy<strong>to</strong>genes strains at low temperature<br />

Cabrita, P., Batista, S., Moes, S., Jenö, P., Trigo, M. J., Boavida Ferreira, R. and Bri<strong>to</strong>, L.<br />

A / O/ 03 Phenotypic and corresponding transcrip<strong>to</strong>mic responses <strong>of</strong> 34<br />

Listeria monocy<strong>to</strong>genes strains in the presence <strong>of</strong><br />

unpro<strong>to</strong>nated organic acids<br />

Lee Chang, K. J., Pinfold, T., Koshy, A. and Bowman, J. P.<br />

A / O/ 04 Internalin pr<strong>of</strong>iling, multilocus sequence typing and 34<br />

virulence assesments suggest evolutionary his<strong>to</strong>ry <strong>of</strong> the<br />

Listeria monocy<strong>to</strong>genes-Listeria innocua clade<br />

Chen, J. and Fang, W.<br />

A / O/ 05 The SOS response <strong>of</strong> Listeria monocy<strong>to</strong>genes is involved 35<br />

in stress resistance, mutagenesis, and bi<strong>of</strong>ilm formation<br />

van der Veen, S. and Abee, T.<br />

A / O/ 06 Clonal diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes, a worldwide perspective 35<br />

Chenal-Francisque, V., Lopez, J., Cantinelli, T., Caro, V., Tran, C., Leclerq, A.,<br />

Lecuit, M. and Brisse, S.<br />

A / O/ 07 Life without a cell wall: Listeria monocy<strong>to</strong>genes L-form cells 36<br />

feature a unique mode <strong>of</strong> division<br />

Briers, Y., Dell’Era, S., Schuppler, M. and Loessner, M. J.<br />

A / O/ 08 Pangenomic analysis <strong>of</strong> Listeria monocy<strong>to</strong>genes 36<br />

Deng, X., Phillippy, A. M., Li, Z., Salzberg, S. L., Tor<strong>to</strong>rello, M. L. and Zhang, W.<br />

A / O/ 09 Evidence for an antiporter-independent glutamate decarboxylase 37<br />

(GAD) system in Listeria monocy<strong>to</strong>genes: Influence <strong>of</strong> growth media<br />

on GAD system activity<br />

Karatzas, K.-A., Brennan, O., Heavin, S. and O’Byrne, C. P.<br />

A / O/ 10 Role <strong>of</strong> Listeria monocy<strong>to</strong>genes tyrosine phosphatases 37<br />

in conferring listeriophage resistance<br />

Paz, R.-N., Eugster, M. R., Zeiman, E., Loessner, M. J. and Calendar, R.<br />

A / O/ 11 Thiolomics – the thiol: disulfide redox metabolism <strong>of</strong> Listeria monocy<strong>to</strong>genes 38<br />

Ondrusch, N., Gopal, S., Fuss, A., Hagen, N., S<strong>to</strong>ll, R., Aharonowitz, Y. and Kreft, J.


A / O/ 12 RNA-structures acting at a distance 38<br />

Johansson, J.<br />

A / O/ 13 Deep RNA sequencing <strong>of</strong> Listeria monocy<strong>to</strong>genes reveals overlapping 39<br />

and extensive stationary phase and sigma B-dependent transcrip<strong>to</strong>mes,<br />

including multiple highly transcribed noncoding RNAs<br />

Oliver, H. F., Orsi, R. H., Ponnala, L., Keich, U., Wang, W., Sun, Q., Cartinhour, S.,<br />

Filiatrault, M. J., Wiedmann, M. and Boor, K. J.<br />

POSTER PRESENTATIONS<br />

A / P/ 00 Compartmentalization <strong>of</strong> IFN-gamma and IL-6 recep<strong>to</strong>rs signalling 62<br />

in Listeria monocy<strong>to</strong>genes phagosomes: immune vesicles<br />

Ramos-Vivas, J., Carrasco-Marin, E., Madrazo-Toca, F., Fernandez-Prie<strong>to</strong>, L.,<br />

Rodriguez-Del Rio, E., Carranza-Cereceda, C. and Alvarez-Dominguez, C.<br />

A / P/ 01 Antimicrobial susceptibility among Listeria monocy<strong>to</strong>genes isolates 63<br />

from non human sources in France over a ten year period<br />

Granier, S. A., Moubarek, C., Colaneri, C., Roussel, S., Courvalin, P. and Brisabois, A.<br />

A / P/ 02 Five homologous small RNAs are involved in the response 63<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes <strong>to</strong> cell wall acting antibiotics<br />

Kiil Nielsen, P. and Kallipolitis, B.<br />

A / P/ 03 Phenotypic analysis <strong>of</strong> selected listerial secretion mutants 64<br />

Halbedel, S., Galander, S. and Flieger, A.<br />

A / P/ 04 Distribution <strong>of</strong> serotypes and pulsotypes <strong>of</strong> L. monocy<strong>to</strong>genes 64<br />

in pig farms (France 2008)<br />

Boscher, E.<br />

A / P/ 05 Comparative phylogenomics <strong>of</strong> Listeria monocy<strong>to</strong>genes reveals 65<br />

an adaptation pr<strong>of</strong>ile<br />

Silveira Nalério, E., Padilha Silva, W., Stabler, R. and Wren, B. W.<br />

A / P/ 06 Serotyping and PFGE patterns <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated 65<br />

from poultry meat<br />

Vasantrao Kurkure, N., Kalorey, D. R., Rodrigues, J., Gunjal, P.1 and Barbuddhe, S. B.<br />

A / P/ 07 Genotypic characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated 66<br />

from fresh leafy vegetables<br />

Warke, S., Kalorey, D. R., Umap, S., Sonegaonkar, A., Patil, V., Kurkure, N V. and Barbuddhe, S. B.<br />

A / P/ 08 Comprehensive appraisal <strong>of</strong> the exoproteome <strong>of</strong> Listeria monocy<strong>to</strong>genes 66<br />

by genomic and proteomic analyses<br />

Desvaux, M., Dumas, E., Chafsey, I., Chambon, C. and Hébraud, M.<br />

A / P/ 09 Investigating differences in lineages <strong>of</strong> Listeria monocy<strong>to</strong>genes 67<br />

using comparative genomics<br />

McIlwham, S., Farber, J. and Pagot<strong>to</strong>, F.<br />

A / P/ 10 Au<strong>to</strong>lysis in Listeria monocy<strong>to</strong>genes – a proteomic approach 67<br />

Pin<strong>to</strong>, E., Marques, N., Andrew, P. W. and Faleiro, M. L.<br />

A / P/ 11 Role <strong>of</strong> flhA, cheR and motA in growth <strong>of</strong> Listeria monocy<strong>to</strong>genes 68<br />

at low temperature<br />

Mattila, M., Lindström, M., Somervuo, P. and Korkeala, H.<br />

A / P/ 12 The effect <strong>of</strong> acetic acid (at pH 5.5) or benzoic acid (at neutral pH) on lipid 68<br />

composition and fluidity <strong>of</strong> Listeria monocy<strong>to</strong>genes membrane<br />

Ioannis, D., Anita, B., Eleni, S. and Mastronicolis, S.<br />

A / P/ 13 Elucidation <strong>of</strong> the responses <strong>to</strong> weak acids in the human pathogen 69<br />

Listeria monocy<strong>to</strong>genes using gene microarrays<br />

O’Byrne, C., Heavin, S. and Morrissey, J.<br />

A / P/ 14 The immunogenic surface protein IspC acts as an N-Acetylglucosaminidase 69<br />

in Listeria monocy<strong>to</strong>genes serotype 4b<br />

Ronholm, J.<br />

A / P/ 15 Listeria monocy<strong>to</strong>genes EGD chitinolytic activity is regulated 70<br />

by carbohydrates but also by the virulence regula<strong>to</strong>ry gene, PrfA<br />

Halberg Larsen, M., Leisner, J. J. and Ingmer, H.<br />

A / P/ 16 Infectious dose curves for guinea pigs challenged with a Listeria 70<br />

monocy<strong>to</strong>genes epidemic clone strain and a strain carrying a<br />

naturally-occurring virulence-attenuating mutation in inlA show a<br />

significant shift in median infectious dose<br />

Nightingale, K., Van Stelten, A., Simpson, J. M., Chen, Y., Scott, V. N., Ross, W. H.,<br />

Whiting, R. C. and Wiedmann, M.<br />

A / P/ 17 MudPIT based proteomic analysis <strong>of</strong> alkaline adapted, environmentally 71<br />

persistent Listeria monocy<strong>to</strong>genes strains<br />

Nilsson, R. E., Ross, T. and Bowman, J. P.<br />

A / P/ 18 RpoN, the alternative sigma fac<strong>to</strong>r, is associated with the growth phase 71<br />

transition and pathogenesis in Listeria monocy<strong>to</strong>genes<br />

Okada, Y., Suzuki, H., Monden, S., Igimi, S. and Okada, N.


A / P/ 19 Cellular lipid fatty acid pattern differences between reference 72<br />

and ice-cream isolate <strong>of</strong> Listeria monocy<strong>to</strong>genes as response <strong>to</strong> cold stress<br />

Anita, B., Ioannis, D. and Mastronicolis, S.<br />

A / P/ 20 Role <strong>of</strong> the dihydroxyace<strong>to</strong>ne metabolism in the resistance 72<br />

<strong>of</strong> Listeria innocua <strong>to</strong> pediocin<br />

Milohanic, E.<br />

A / P/ 21 Glucose transport system in Listeria monocy<strong>to</strong>genes 73<br />

and their impact on virulence gene expression<br />

Moussan Ake, F.<br />

A / P/ 22 Antimicrobial susceptibilities <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated in Japan 73<br />

Monden, S., Okutani, A., Suzuki, H., Asakura, H., Nakama, A., Igimi, S., Okada, Y. and Maruyama, T.<br />

A / P/ 23 Influence <strong>of</strong> sub-lethal concentrations <strong>of</strong> disinfectants 74<br />

on Listeria monocy<strong>to</strong>genes adhesion and invasion in Caco-2 cells<br />

Gaedt Kastbjerg, V., Halberg Larsen, M., Ingmer, H. and Gram, L.<br />

A / P/ 24 The SOS response in Listeria monocy<strong>to</strong>genes – a stress survival mechanism 74<br />

Kiil Nielsen, P., Zahle Andersen, A. and Haahr Kallipolitis, B.<br />

A / P/ 25 Acid shock triggers heavy metal de<strong>to</strong>xification in Listeria monocy<strong>to</strong>genes 75<br />

Müller, S., Neuhaus, K. and Sc<strong>here</strong>r, S.<br />

A / P/ 26 Listeria monocy<strong>to</strong>genes mutants defective in growth at 5 °C 75<br />

and in high salt environment<br />

Burall, L., Laksanalamai, P. and Datta, A.<br />

A / P/ 27 Revival <strong>of</strong> 5,000 Listeria strains from Seeliger’s his<strong>to</strong>rical collection 76<br />

with a semi-au<strong>to</strong>mated microbiological pipeline<br />

Haase, J., H<strong>of</strong>, H. and Achtman, M.<br />

A / P/ 28 Two point mutations are responsible for the lack <strong>of</strong> glycosidic substition 76<br />

in cell wall teichoic acids in Listeria monocy<strong>to</strong>genes serovar “7”<br />

Eugster, M. R., Huwiler, S., Morax, L. and Loessner, M. J.<br />

A / P/ 29 High-throughput genome sequencing <strong>of</strong> two Listeria monocy<strong>to</strong>genes 77<br />

clinical isolates during a large foodborne outbreak<br />

Gilmour, M., Graham, M., Van Domselaar, G., Tyler, S., Kent, H., Trout-Yakel, K.,<br />

Larios, O., Allen, V., Lee, B., Nadon, C. and Kearney, A.<br />

A / P/ 30 Expression <strong>of</strong> antimicrobial activity in food 77<br />

and clinical Listeria monocy<strong>to</strong>genes isolates<br />

Barbosa, J., Ferreira, V., Borges, S., Azevedo, I., Magalhães, R., San<strong>to</strong>s, I, Almeida, G. and Teixeira, P.<br />

A / P/ 31 The Listeria monocy<strong>to</strong>genes sigma B and sigma H regulons overlap, 78<br />

but only sigma B appears <strong>to</strong> be important for survival <strong>of</strong> acid,<br />

alkaline and oxidative stress<br />

Chaturongakul, S., Raengpradub, S., Wiedmann, M. and Boor, K. J.<br />

A / P/ 32 Virulence gene expression in Listeria monocy<strong>to</strong>genes strains isolated 78<br />

from different sources<br />

Alessandria, V., Rantsiou, K. and Cocolin, L.<br />

A / P/ 33 Differentiation <strong>of</strong> Listeria monocy<strong>to</strong>genes, Listeria innocua 79<br />

and Listeria marthii, a novel Listeria species isolated<br />

from the natural environment, Finger Lakes National Forest<br />

Graves, L. M., Helsel, L. O., Steigerwalt, A. G., Morey, R. E., Daneshvar, M. I., Ro<strong>of</strong>, S. E.,<br />

Orsi, R. H., Fortes, E. D., Milillo, S. R., den Bakker, H. C., Wiedmann, M.,<br />

Swaminathan, B. and Sauders, B. D.<br />

A / P/ 34 Differed roles <strong>of</strong> L,D-carboxypeptidases encoded by lmo0028 79<br />

and lmo1638 genes<br />

Yurov, D., Varfolomeev, A., Kaminskaya, A. and Ermolaeva, S.<br />

A / P/ 35 Antimicrobial susceptibilities <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated 80<br />

from retail beef, pork and poultry in Japan<br />

Ida, M., Shimojima, Y., Kaneko, S., Higuchi, Y., Nakama, A. and Kai, A.<br />

A / P/ 36 Genetic basis <strong>of</strong> two low pathogenic L. monocy<strong>to</strong>genes strains 80<br />

with apparent phospholipase C activity<br />

Jiang, L., Bai, F., Chen, J., and Fang, W.<br />

A / P/ 37 Molecular genotyping and antimicrobial resistance 81<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes from foods and the environment<br />

Parisi, A., Miccolupo, A., Fraccalvieri, R., La<strong>to</strong>rre, L., Normanno, G. and Santagada, G.<br />

A / P/ 38 Virulence transcrip<strong>to</strong>me analysis <strong>of</strong> Listeria monocy<strong>to</strong>genes 81<br />

by application <strong>of</strong> microarrays in vitro and in situ<br />

Rantsiou, K., Alessandria, V. and Cocolin, L.<br />

A / P/ 39 Effects <strong>of</strong> growth conditions on surface properties <strong>of</strong> Listeria; 82<br />

a proposed role for AI-2<br />

Wong, H. T. L., Nwaiwu, O. and Rees, C. E. D.<br />

A/P/40 Stress behaviour <strong>of</strong> a Listeria monocy<strong>to</strong>genes 568 Lmo1634 transposon mutant 82<br />

Truelstrup Hansen, L., Holman, D. B. and Ells, T. C.


A/P/41 Investigation <strong>of</strong> the conditions that trigger the activation 83<br />

<strong>of</strong> the alternative sigma fac<strong>to</strong>r σB in Listeria monocy<strong>to</strong>genes<br />

Utratna, M., Shaw, I. and O’Byrne, C.<br />

A/P/42 Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes 1/2a, 1/2b, 1/2c and 4b 83<br />

by amplified fragment length polymorphism and evaluation<br />

<strong>of</strong> their geographical distributions in Portugal<br />

Maia, C. H., Goulão, M. M., San<strong>to</strong>s, M. I., Ferreira, M. A. S. S. and Pintado, C. M. B. S.<br />

A/P/43 Global analysis <strong>of</strong> the Listeria monocy<strong>to</strong>genes surface proteins 84<br />

<strong>of</strong> the LPXTG family<br />

Botello-Morte, L., Calvo, E., Mariscotti, J., D’Orazio, V., García-del Portillo, F. and Pucciarelli, M. G.<br />

A/P/44 Antimicrobial susceptibility <strong>of</strong> Listeria monocy<strong>to</strong>genes strains 84<br />

derived from food and food-processing bakery plant<br />

Eusébio, C., Carneiro, L., San<strong>to</strong>s, I., Magalhães, R., Almeida, G., Silva, J. and Teixeira, P.<br />

A/P/45 A physiological study <strong>to</strong> purpose a new formal method 85<br />

<strong>to</strong> obtain L. monocy<strong>to</strong>genes cells adapted <strong>to</strong> BAC<br />

Saá Ibusquiza, P., Cabo, M. L., Herrera, J. J. R., Vázquez, D., Carrera, S. and Eiriz, E.<br />

A/P/46 Characterization <strong>of</strong> a RNA-helicase in the human pathogen 85<br />

Listeria monocy<strong>to</strong>genes<br />

Netterling, S. and Johansson, J.<br />

A/P/47 Listeria monocy<strong>to</strong>genes agr system: Quorum sensing, or maybe not? 162<br />

Garmyn, D., Révelin, C. and Piveteau, P.<br />

AREA //B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

PLENARY LECTURES<br />

B / PL/ 03 How Listeria monocy<strong>to</strong>genes breaches host barriers 23<br />

Lecuit, M.<br />

B / PL/ 04 Secretion <strong>of</strong> a novel L. monocy<strong>to</strong>genes cyclic dinucleotide 23<br />

in<strong>to</strong> the cy<strong>to</strong>sol <strong>of</strong> infected host cells activates an innate immune pathway<br />

Portnoy, D. A.<br />

B / PL/ 05 Diagnosis and clinical management <strong>of</strong> listeriosis in ruminants and camelids 24<br />

Poulsen, K. P.<br />

B / PL/ 06 Listeriolysin S – a second haemolysin with a role 24<br />

in the virulence <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Hill, C.<br />

ORAL PRESENTATIONS<br />

B / O/ 14 Probiotics reduce Listeria monocy<strong>to</strong>genes-induced tissue invasion 40<br />

and stillbirths in pregnant guinea pigs<br />

Smith, M. A., Agyekum, K. and Williams, D.<br />

B / O/ 15 Listeriolysin O favors Listeria monocy<strong>to</strong>genes growth 40<br />

in co-culture with the ciliate Tetrahymena pyriformis<br />

Ermolaeva, S. and Pushkareva, V.<br />

B / O/ 16 A<strong>to</strong>py is a risk fac<strong>to</strong>r for listeriosis 41<br />

Kawamo<strong>to</strong>, K., Matsubara, S., Da Silva, M. and Makino, S.-I.<br />

B / O/ 17 Cancer immunotherapy using novel Listeria monocy<strong>to</strong>genes-bacterial 41<br />

vec<strong>to</strong>rs <strong>to</strong> target the vasculature <strong>of</strong> progressive tumors<br />

Paterson, Y., Seavey, M. M., Maciag, P. C. and Sewell, D.<br />

B / O/ 18 The intracellular carbon metabolism <strong>of</strong> Listeria monocy<strong>to</strong>genes 42<br />

in comparison <strong>to</strong> that <strong>of</strong> other intracellular bacterial pathogens<br />

replicating in mammalian host cells<br />

Gotz, A., Eylert, E., S<strong>to</strong>ll, R., Eisenreich, W., and Goebel, W.<br />

B / O/ 19 LIMP2 links late phagosomal trafficking with the onset <strong>of</strong> the innate 42<br />

immune response <strong>to</strong> Listeria monocy<strong>to</strong>genes: a role in macrophage activation<br />

Fernandez-Prie<strong>to</strong>, L., Carrasco-Marin, E., Madrazo-Toca, F., Rodriguez-Del Rio, E.,<br />

Carranza-Cereceda, C. and Alvarez-Dominguez, C.<br />

B / O/ 20 The tetraspanin CD81 is required for entry <strong>of</strong> Listeria monocy<strong>to</strong>genes 43<br />

in mammalian cells<br />

Tham, T. N., Gouin, E., Rubinstein, E., Boucheix, C., Cossart, P. and Pizarro-Cerdá, J.<br />

B / O/ 21 Listeria innate immune evasion by peptidoglycan modification 43<br />

Aubry, C.<br />

B / O/ 22 Constitutive activation <strong>of</strong> the central virulence transcriptional regula<strong>to</strong>r 44<br />

PrfA enhances Listeria monocy<strong>to</strong>genes pathogenesis<br />

but reduces bacterial fitness outside <strong>of</strong> the host<br />

Freitag, N. E. and Bruno, J. C.<br />

B / O/ 23 The lvfH gene <strong>of</strong> Listeria monocy<strong>to</strong>genes encodes a novel 44<br />

virulence fac<strong>to</strong>r highly activated during infection<br />

Carvalho, F., Camejo, A., Ferreira, P., Sousa, S. and Cabanes, D.


POSTER PRESENTATIONS<br />

B / P/ 47 A Listeria monocy<strong>to</strong>genes strain is still virulent despite non-functional 86<br />

major virulence genes: Optical Mapping shows a potential mechanism<br />

Roche, S. M., Grépinet, O., Corde, Y., Teixeira, A. P., Kerouan<strong>to</strong>n, A., Témoin, S.,<br />

Mereghetti, L., Brisabois, A. and Velge, P.<br />

B / P/ 48 Protein expression <strong>of</strong> lineage I, II, and III Listeria monocy<strong>to</strong>genes 86<br />

strains in murine macrophages<br />

Donaldson, J. R., Nanduri, B., Pittman, J. R., Burgess, S. C. and Lawrence, M. L.<br />

B / P/ 49 Prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in bovine mastitic milk samples: 87<br />

Possible source <strong>of</strong> food borne infection<br />

Kalorey, D. R., Warke, S., Kurkure, N. V. and Barbuddhe, S. B.<br />

B / P/ 50 Agr-dependent peptide sensing in L. monocy<strong>to</strong>genes – Effects on 87<br />

bi<strong>of</strong>ilm formation, virulence and global gene expression<br />

Waidmann, M. S., Monk, I. R., Auchter, M., Preising, N. P., Hill, C. and Riedel, C. U.<br />

B / P/ 51 agrD-deletion affects InlA- and InlB-regulation via 88<br />

a temperature-dependent and -independent mechanism<br />

Waidmann, M. S., Monk, I. R. and Riedel, C. U.<br />

B / P/ 52 Bovine cranial nerve Schwann cells express E-cadherin, 88<br />

a candidate key-player in the brainstem invasion<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes in cattle<br />

Madarame, H., Seuberlich, T., Vandevelde, M., Zurbriggen, A., and Oevermann, A.<br />

B / P/ 53 Pathogenic potential <strong>of</strong> Listeria monocy<strong>to</strong>genes isolates 89<br />

from New Zealand seafood premises: implications for control<br />

Durante Cruz, C. and Fletcher, G.<br />

B / P/ 54 Copper homeostasis and virulence in Listeria monocy<strong>to</strong>genes 89<br />

David, C., Schuler, S., Glenn, S., Jen, C., Andrew, P. and Roberts, I. S.<br />

B / P/ 55 Pattern <strong>of</strong> cy<strong>to</strong>kine production during murine listeriosis 90<br />

Dussurget, O.<br />

B / P/ 56 Analysis <strong>of</strong> the post-translocation chaperone PrsA2 and its unique 90<br />

role in facilitating Listeria monocy<strong>to</strong>genes pathogenesis<br />

Alonzo, F. and Freitag, N.<br />

B / P/ 57 CtaP is a multifunctional cysteine-transport associated protein 91<br />

required for Listeria monocy<strong>to</strong>genes pathogenesis<br />

Xayarath, B.<br />

B / P/ 58 Enzymatic activity <strong>of</strong> the metalloprotease <strong>of</strong> Listeria is regulated by pH 91<br />

Forster, B. M., Pavinski Bitar, A., Slepkov, E. R. and Marquis, H.<br />

B / P/ 59 Identification <strong>of</strong> propeptide residues regulating the compartmentalization, 92<br />

maturation, and activity <strong>of</strong> the broad-range phospholipase C<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Slepkov, E. R., Pavinski Bitar, A. and Marquis, H.<br />

B / P/ 60 A mouse model <strong>of</strong> fe<strong>to</strong>placental Listeria monocy<strong>to</strong>genes 92<br />

infection and abortion<br />

Poulsen, K. P., Faith, N., Laura Knoll, L. and Czuprynski, C.<br />

B / P/ 61 Listeria infection <strong>of</strong> the insect model system Galleria mellonella 93<br />

Joyce, S. A. and Gahan, C. G.<br />

B / P/ 62 Construction <strong>of</strong> a murinised Listeria monocy<strong>to</strong>genes H7858 (4b) 93<br />

strain for improved murine infection<br />

Cummins, J. and Gahan, C.<br />

B / P/ 63 The role <strong>of</strong> a phosphoinositide phosphatase in the intracellular survival 94<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Wang, J., Corbett, D. and Roberts, I. S.<br />

B / P/ 64 Virulence gene expression in Listeria monocy<strong>to</strong>genes strains isolated 94<br />

from different sources<br />

Alessandria, V.<br />

B / P/ 65 Targeted Signature-tagged mutagenesis for phenotype screening 95<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes mutants<br />

Henriques, A., Carvalho, F. and Cabanes, D.<br />

B / P/ 66 Listeria monocy<strong>to</strong>genes cellular infection triggers 95<br />

tyrosine-phosphorylation <strong>of</strong> Myosin IIA, a new protein<br />

involved in invasion<br />

Almeida, M. T., Cabanes, D. and Sousa, S.<br />

B / P/ 67 Invasion pr<strong>of</strong>ile <strong>of</strong> Listeria monocy<strong>to</strong>genes strains involved 96<br />

in invasive and gastroenteritis listeriosis outbreaks<br />

Laksanalamai, P., Sahu, S. and Datta, A.<br />

B / P/ 68 Molecular characterization <strong>of</strong> the Vip-Gp96 interaction 96<br />

Martins, M., Cabanes, D. and Sousa, S.<br />

B / P/ 69 Investigation <strong>of</strong> the molecular mechanisms by which 97<br />

Listeria monocy<strong>to</strong>genes grows in the mammalian gall bladder<br />

Dowd, G., Joyce, S., Casey, P. G., Hill, C. and Gahan, C. G.


B / P/ 70 Role <strong>of</strong> cadmium efflux system in Listeria monocy<strong>to</strong>genes virulence 97<br />

Camejo, A. and Cabanes, D.<br />

B / P/ 71 Investigation <strong>of</strong> chitinase as a potential virulence fac<strong>to</strong>r 98<br />

in Listeria monocy<strong>to</strong>genes<br />

Chaudhuri, S.<br />

B / P/ 72 Sub-lethal concentrations <strong>of</strong> common disinfectants do not 98<br />

influence survival and growth <strong>of</strong> Listeria monocy<strong>to</strong>genes in whole blood<br />

Holch, A., Gaedt Kastbjerg, V. and Gram, L.<br />

B / P/ 73 Internalin LRR domain variability in Listeria monocy<strong>to</strong>genes 99<br />

isolated from different hosts<br />

Zaytseva, E. A., Ermolaeva, S. A. and Somov, G. P.<br />

B / P/ 74 Reduced virulence <strong>of</strong> an adenylosuccinate lyase transposon mutant 99<br />

<strong>of</strong> a serotype 4b strain <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Faith, N. G., Kim, J.-W., Kathariou, S., Sahaghian, R. and Luchansky, J. B.<br />

B / P/ 75 Manifestations and outcome <strong>of</strong> listeriosis in adult patients 100<br />

Fernández Guerrero, M L., Mancebo Plaza, B., Torres, R., Górgolas, M. and Jusdado, J. J.<br />

B / P/ 76 Model for human Listeriosis: in vivo moni<strong>to</strong>ring <strong>of</strong> orally infected mice 100<br />

using bioluminescent Listeria monocy<strong>to</strong>genes<br />

Bergmann, S., Lengeling, A., Pasche, B. and Schughart, K.<br />

B / P/ 77 Galleria mellonella as model system <strong>to</strong> study Listeria species-specific 101<br />

and Listeria monocy<strong>to</strong>genes serotype-specific pathogenesis<br />

Mraheil, M. A., Krishnendu, M., Hain, T. and Chakraborty, T.<br />

B / P/ 78 Listeria monocy<strong>to</strong>genes ActA is a key player in evading au<strong>to</strong>phagic recognition 101<br />

Pillich, H., Loose, M., Hain, T. and Chakraborty, T.<br />

B / P/ 79 Non-haemolytic and hypovirulent Listeria monocy<strong>to</strong>genes became 102<br />

haemolytic and virulent after passage through mice<br />

Secic, I., Lindbäck, T. and Rørvik, L. M.<br />

B / P/ 80 Mutants <strong>of</strong> Listeria monocy<strong>to</strong>genes (Lm) resistant <strong>to</strong> the polycationic 102<br />

peptide protamine appear <strong>to</strong> be attenuated for virulence<br />

Schlech, W.<br />

B / P/ 81 The role <strong>of</strong> plasmacy<strong>to</strong>id dendritic cells in the course 103<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes infection<br />

Solodova, E., Lienenklaus, S., Jablonska, J. and Weiss, S.<br />

B / P/ 182 Oxygen restriction increases the infection potential <strong>of</strong> 103<br />

Listeria monocy<strong>to</strong>genes – a transcriptional analysis<br />

Andersen, J. B., Bergstrøm, A., Knudsen, G., Bak Christensen, B., Ebersbach, T.,<br />

Boye, M. and Rask Licht, T.<br />

AREA //C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

and listeriosis<br />

PLENARY LECTURES<br />

C / PL/ 07 Update on clinical aspects <strong>of</strong> listeriosis: What’s new? 25<br />

Schlech, W. F.<br />

C / PL/ 08 A Canadian outbreak <strong>of</strong> listeriosis due <strong>to</strong> deli-meat: 25<br />

Driving change in the food safety system<br />

Farber, J. M., Pagot<strong>to</strong>, F., Gilmour, M., Nadon, C., Savelli, C., and MacDonald, D.<br />

C / PL/ 09 Listeria monocy<strong>to</strong>genes phagocytic strategy 26<br />

Alvarez-Dominguez, C.<br />

ORAL PRESENTATIONS<br />

C / O/ 24 Human listeriosis due <strong>to</strong> Listeria ivanovii 45<br />

Guillet, C., Join-Lambert, O., Le Monnier, A., Leclercq, A., Mamzer-Bruneel, M. F.,<br />

Bielecka, M. K., Scortti, M., Disson, O., Vazquez-Boland, J., Lortholary, O. and Lecuit, M.<br />

C / O/ 25 Foodborne Listeriosis in India: An update 45<br />

Barbuddhe, S. B., Malik, S. V. S., Ashok Kumar, J., Kalorey, D. R., Kurkure, N. V.,<br />

Rawool, D. B., Swain, B. K., Korikanthimath, V. S., Chakraborty, T.<br />

C / O/ 26 Human listeriosis and co-morbidities in England, 1999 <strong>to</strong> 2008: 46<br />

quantifying the risk<br />

Mook, P., Grant, K., O’Brien, S. J. and Gillespie, I.<br />

C / O/ 27 Risk fac<strong>to</strong>rs for death in Listeria monocy<strong>to</strong>genes infection, 46<br />

England and Wales, 1990 <strong>to</strong> 2008<br />

Mook, P., Grant, K. and Gillespie, I.<br />

C / O/ 28 A discrete event model <strong>to</strong> track Listeria monocy<strong>to</strong>genes 47<br />

in the retail environment<br />

Pouillot, R. and Gallagher, D.


C / O/ 29 Clinical and epidemiological aspects <strong>of</strong> listeriosis in Israel 47<br />

Hershko-Klement, A., Eliav, H., Valinsky, L., Schechner, V., Braun, E., Paitan, Y.,<br />

Block, C. S. and Nir-Paz, R.<br />

C / O/ 30 Human isolates <strong>of</strong> Listeria monocy<strong>to</strong>genes during half a century in Sweden 48<br />

Lopez-Valladares, G., Danielsson-Tham, M.-L., Vishal Singh, P. and Tham, W.<br />

C / O/ 31 Estimated incubation periods for listeriosis vary according 48<br />

<strong>to</strong> clinical form <strong>of</strong> disease<br />

Goulet, V., King, L., Vaillant, V. and De Valk, H.<br />

C / O/ 32 Listeria monocy<strong>to</strong>genes in food and animals in the European Union in 2008 49<br />

da Silva Felício, M. T., Rizzi, V., Boelaert, F. and Makela, P.<br />

C / O/ 33 Listeria monocy<strong>to</strong>genes infection in the over 60s in England 49<br />

between 2005 and 2008: a retrospective case-control study utilising<br />

market research panel data<br />

Gillespie, I. A., Mook, P., Little, C. L. and Grant, K.<br />

C / O/ 34 Better and faster typing, MLVA – shall we play <strong>to</strong>gether? 50<br />

Larsson, J. T., Roussel, S. and Moller Nielsen, E.<br />

C / O/ 35 comK prophage junction fragments in Listeria monocy<strong>to</strong>genes contain 50<br />

SNPs that differentiate subclones <strong>of</strong> ECII and ECIII that are unique<br />

<strong>to</strong> individual meat and poultry processing plants in the U.S.<br />

Knabel, S. J.<br />

C / O/ 36 Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes measured by multiple-locus 51<br />

variable-number tandem repeat analysis (MLVA)<br />

Hyytia-Trees, E., Sabol, A., Graves, L. and Ribot, E.<br />

POSTER PRESENTATIONS<br />

C / P/ 82 Semi-au<strong>to</strong>mated repetitive sequenced-based PCR compared 104<br />

<strong>to</strong> pulsed field gel electrophoresis for Listeria monocy<strong>to</strong>genes sub-typing<br />

Roussel, S., Félix, B., Vignaud, M.-L., Tam Dao, T., Marault, M. and Brisabois, A.<br />

C / P/ 83 Listeriosis: a frequent cause <strong>of</strong> fatal encephalitis in France 104<br />

with high case fatality<br />

Mailles, A., Vaillant, V., Lecuit, M. and Stahl, J.-P.<br />

C / P/ 84 Listeria monocy<strong>to</strong>genes: identification and subtyping 105<br />

Favretti, M.<br />

C / P/ 85 Virulotyping <strong>of</strong> Listeria monocy<strong>to</strong>genes by high resolution melt analysis 105<br />

Amar, C. F., Tamburro, M., Dear, P. and Grant, K.<br />

C / P/ 86 Risk fac<strong>to</strong>rs for nonperinatal listeriosis mortality in Los Angeles County, 106<br />

California, 1992–2004<br />

Guevara, R. E., Mascola, L. and Sorvillo, F.<br />

C / P/ 87 Molecular typing <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated from ovine sausage 106<br />

Nives, M. R., Mele, P., Parisi, A., La<strong>to</strong>rre, L., Virgilio, S. and Tola, S.<br />

C / P/ 88 A novel phage-PCR assay for the rapid detection <strong>of</strong> viable 107<br />

Listeria monocy<strong>to</strong>genes in food within 24 h<br />

Elemam, M. M.<br />

C / P/ 89 Perinatal listeriosis in Los Angeles County, California, 1992–2004 107<br />

Guevara, R. E. and Mascola, L.<br />

C / P/ 90 Antimicrobial resistance <strong>of</strong> Listeria monocy<strong>to</strong>genes human strains 108<br />

isolated since 1926 in France<br />

Morvan, C., Moubareck, A., Leclercq, M., Herve-Bazin, S., Bremont, M.,<br />

Lecuit, P. and Le Monnier Courvalin, A.<br />

C / P/ 91 Today and <strong>to</strong>morrow: The molecular epidemiology <strong>of</strong> listeriosis in the UK 108<br />

Grant, K., Amar, C., Ma<strong>to</strong>s, J., Mook, P., Little, C. and Gillespie, I.<br />

C / P/ 92 Use <strong>of</strong> Fluorescent Amplified Fragment Length Polymorphism 109<br />

for improved typing <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Ma<strong>to</strong>s, J., Amar, C., Desai, M., Cross, L. and Grant, K.<br />

C / P/ 93 Towards an application for field samples <strong>of</strong> a multipathogen platform 109<br />

for molecular detection <strong>of</strong> raw milk pathogens<br />

Omiccioli, E, Amagliani, G., Brandi, G., Tonucci, F., Foglini, M. and Magnani, M.<br />

C / P/ 94 Detection and recovery <strong>of</strong> Listeria species from stainless steel 110<br />

Boone, R., Iugovaz, I, Trottier, Y.-L. and Pagot<strong>to</strong>, F.<br />

C / P/ 95 Natural carriage <strong>of</strong> Listeria in fresh-water fish in Russia 110<br />

Egorova, I., Voronin, M., Selyaninov, Y. and Kolbasov, D.<br />

C / P/ 96 Prevalence <strong>of</strong> Listeria in wild fauna in Russia 111<br />

Egorova, I., Fertikov, V. and Kolbasov, D.<br />

C / P/ 97 Quantitative assessment <strong>of</strong> the exposure <strong>to</strong> Listeria monocy<strong>to</strong>genes 111<br />

from s<strong>of</strong>t-ripened cheese consumption in North America:<br />

a joint FDA/Health Canada project<br />

Gendel, S., Pouillot, R., Murray, C., Farber, J., Ross, W., Couture, H. and Jean, A.


C / P/ 98 Human listeriosis in England, 2001 – 2007: association with 112<br />

neighbourhood deprivation<br />

Gillespie, I. A., Mook, P., Little, C. L., Grant, K. and McLauchlin, J.<br />

C / P/ 99 Characterisation <strong>of</strong> antibodies for use in an antibody-based sensor 112<br />

for on-site detection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

Gilmartin, N., Hearty, S. and O’ Kennedy, R.<br />

C / P/ 100 Food Investigation <strong>of</strong> sporadic cases <strong>of</strong> neuroinvasive listeriosis 113<br />

Goulet, V., Leclercq, A., Laurent, E., King, L., Dusch, V., Salem, S., Vaillant, V.,<br />

Chenal-Francisque, V., de Valk, H. and Pihier, N.<br />

C / P/ 101 Evaluation <strong>of</strong> the antimicrobial activity <strong>of</strong> Vaccinium myrtillus, 113<br />

Prunus domestica and Myrtus communis against Listeria monocy<strong>to</strong>genes<br />

Serio, A., Di Pasquale, F. and Paparella, A.<br />

C / P/ 102 High throughput quantitative detection <strong>of</strong> Listeria monocy<strong>to</strong>genes 114<br />

in food by RealTime PCR<br />

Cammà, C., Ancora, M., Rizzi, V., Sperandii, A., Prencipe, V. and Migliorati, G.<br />

C / P/ 103 Bibliographic study concerning procedures for preparing 114<br />

environmental samples for analyses, regarding <strong>to</strong> L. monocy<strong>to</strong>genes<br />

Barre, L., Carpentier, B. and Gnanou Besse, N.<br />

C / P/ 104 Persistent L. monocy<strong>to</strong>genes isolates from Austrian and Irish dairies 115<br />

show the same phenotypic and genetic background<br />

Stessl, B.<br />

C / P/ 105 Epidemiological data on listeriosis in Portugal: 2003 – 2008 115<br />

Almeida, G, Magalhães, R., Hogg, T. and Teixeira, P.<br />

C / P/ 106 Diversity among Listeria monocy<strong>to</strong>genes isolated from humans 116<br />

Kalekar, S., Rodrigues, J., D’Costa, D., Malik, S. V. S., Kalorey, D. R.,<br />

Chakraborty, T. and Barbuddhe, S. B.<br />

C / P/ 107 Characterization <strong>of</strong> Listeria isolated from seafood 116<br />

Rodrigues, J., Kalekar, S., Bhosle, S. N., Doijad, S. and Barbuddhe, S. B.<br />

C / P/ 108 Characterization <strong>of</strong> Listeria species isolated from milk 117<br />

D’Costa, D., Bhosle, S. N., Dhuri, R. B., Kalekar, S., Rodrigues, J., Doijad, S. P. and Barbuddhe, S. B.<br />

C / P/ 109 Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in chicken production chain 117<br />

in Thailand<br />

Kanarat, S., Nijthavorn, N. and Sukhapesna, J.<br />

C/P/110 Listeria monocy<strong>to</strong>genes in Ireland – Epidemiology and Molecular Typing 118<br />

Cormican, M. G., DeLappe, N., McKeown, P. and Garvey, P.<br />

C / P/111 Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated 118<br />

from human cases <strong>of</strong> listeriosis occurred in Portugal in 2008<br />

Magalhães, R., Barbosa, J., San<strong>to</strong>s, I., Almeida, G. and Teixeira, P.<br />

C / P/112 Post-processing environmental contamination <strong>of</strong> surface-ripened 119<br />

s<strong>of</strong>t cheese during affinage<br />

D’Amico, D. and Donnelly, C.<br />

C / P/113 Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated 119<br />

from Portuguese cheeses<br />

Almeida, G., Magalhães, R., San<strong>to</strong>s, I., Barbosa, J., Hogg, T. and Teixeira, P.<br />

C / P/114 Prevalence <strong>of</strong> Listeria spp. in retail raw ground beef in Izmir, Turkey: 120<br />

A comparison <strong>of</strong> standard cultural method and Fluorescent in situ<br />

Hybridization (FISH) technique for detection<br />

Handan Baysal, A.<br />

C / P/115 Using ListexP100 for Listeria monocy<strong>to</strong>genes detection in foods 120<br />

Flores Lopes, J., Ferreira Leite, I., Azeredo, J., Gibbs, P. and Teixeira, P.<br />

C / P/116 Mapping <strong>of</strong> molecular pr<strong>of</strong>iles associated <strong>to</strong> Listeria monocy<strong>to</strong>genes 121<br />

food isolates circulating in Italy<br />

De Cesare, A., Parisi, A., La<strong>to</strong>rre, L. and Manfreda, G.<br />

C / P/117 Zoonotic aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes isolates 121<br />

from zebu dairy animals<br />

Parihar, V. S. , Barbuddhe, S. B., Kalorey, D. R., Kotwal, S.,Danielsson Tham, M.-L. and Tham, W.<br />

C / P/118 Performance <strong>of</strong> ALOA and Palcam agars for detection <strong>of</strong> 122<br />

Listeria monocy<strong>to</strong>genes in naturally contaminated raw meat products<br />

Grava<strong>to</strong> Rowlands, R. E., Asturiano Ris<strong>to</strong>ri, C. A., Geraldes Martins, C.,<br />

Jakabi, M. and Gombossy de Melo Franco, B. D.<br />

C / P/119 Comparison <strong>of</strong> MOPS-BLEB and Fraser as secondary enrichment broths 122<br />

for Listeria monocy<strong>to</strong>genes<br />

Upham, J., Huszczynski, G., Mosher, M., Borza, A., Dorey, M., Bosley, J., Hara, K.,<br />

Mutanda, C., Liu, J., Byrne, B. and Douey, D.<br />

C / P/120 Comparison <strong>of</strong> UVM, Palcam and Oxoid Novel Enrichment (ONE) 123<br />

broth as primary enrichment broths for Listeria monocy<strong>to</strong>genes<br />

Upham, J., Mosher, M., Borza, A., Huszczynski, G., Dorey, M., Eloranta, K. and Douey, D.<br />

C / P/121 Isolation and characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes 123<br />

from asazuke (Japanese light pickles)<br />

Maklon, K., Kusumo<strong>to</strong>, A., Makino, S.-I. and Kawamo<strong>to</strong>, K.


C / P/122 Grouping <strong>of</strong> human Listeria monocy<strong>to</strong>genes isolates 124<br />

Lopez-Valladares, G., Danielsson-Tham, M.-L. and Tham, W.<br />

C / P/123 Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated 124<br />

from food and environmental samples<br />

Nucera, D. M., Lomonaco, S., Manila Bianchi, D., Decastelli, L., Grassi, M. A. and Civera, T.<br />

C / P/124 Development <strong>of</strong> a multiplex SNP-typing method <strong>to</strong> identify 125<br />

the four epidemic clones <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Lomonaco, S., Civera, T., Dalmasso, A., Knabel, S. J. and Bottero, M. T.<br />

C / P/125 Listeria monocy<strong>to</strong>genes incidents reported <strong>to</strong> the UK 125<br />

Food Standards Agency from 2000 <strong>to</strong> 2008<br />

Aish, J.<br />

C / P/183 Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes in broiler flocks 126<br />

Courtillon, C. Toquin, M.-T., Le Nôtre, Y., Fravalo, P. and Mansour Chemaly, M.<br />

C / P/184 Tracing Listeria monocy<strong>to</strong>genes contaminations throughout 126<br />

the processing chain <strong>of</strong> a typical Italian pork meat product<br />

using Pulsed Field Gel Electrophoresis (PFGE) characterization<br />

Annunziata Prencipe, V., Acciari, V., Torresi, M., Migliorati, G.,<br />

Marfoglia, C. and Valentina Rizzi, V.<br />

C / P/185 Production and validation <strong>of</strong> capture-ELISA kit based on monoclonal 127<br />

antibodies specific for Listeria monocy<strong>to</strong>genes in foodstuffs<br />

Portanti, O., Di Febo, T., Luciani, M., Pompilii, C., Armillotta, G., Principe, V.,<br />

Lelli, R. and Semprini, P.<br />

C / P/186 Production <strong>of</strong> a reference material for microbiological tests 127<br />

containing Listeria innocua<br />

Pomilio, F., Ricci, L, Di Giannatale, E., Semprini, P., Candeloro, L. and Migliorati, G.<br />

AREA //D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

PLENARY LECTURES<br />

D / PL/ 10 Risk assessment using the microbiological criteria: arguments for 27<br />

zero <strong>to</strong>lerance (USDA) other viewpoint (Europe):<br />

Risk-based microbiological criteria for Listeria monocy<strong>to</strong>genes<br />

in RTE foods<br />

Luber, P.<br />

D / PL/ 11 USDA regula<strong>to</strong>ry approach and considerations for the control 27<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Engeljohn, D.<br />

D / PL/ 12 Listeria monocy<strong>to</strong>genes, an emergent pathogen in Chile and Latin America 28<br />

Hormazábal, J. C.<br />

ORAL PRESENTATIONS<br />

D / O/ 37 Time Temperature Indica<strong>to</strong>rs can be used as an effective 52<br />

Risk Management Tool for Listeria monocy<strong>to</strong>genes in Ready-To-Eat foods<br />

Koutsoumanis, K., Vaikousi, H. and Costas, B.<br />

D / O/ 38 Safety <strong>of</strong> Salad Leaves and Herbs 52<br />

Garland, C. D. and Clark, A.<br />

D / O/ 39 Environmental fac<strong>to</strong>rs affect adhesion <strong>of</strong> Listeria monocy<strong>to</strong>genes 53<br />

<strong>to</strong> inert surfaces through flagellum expression<br />

Tresse, O.<br />

D / O/ 40 Shelf-life labora<strong>to</strong>ry durability and challenge studies for Listeria 53<br />

monocy<strong>to</strong>genes in ready-<strong>to</strong>-eat foods: a presentation <strong>of</strong> the<br />

European technical guidance document intended for labora<strong>to</strong>ries<br />

Beaufort, A., Bergis, H., Cornue, M. and Lardeux, A.-L.<br />

D / O/ 41 Successful strategies against Listeria monocy<strong>to</strong>genes in Switzerland 54<br />

Imh<strong>of</strong>, R.<br />

D / O/ 42 Absence <strong>of</strong> Listeria monocy<strong>to</strong>genes growth during raw milk cheesemaking: 54<br />

a modelling approach<br />

Jordan, K., Schvartzman, S. Butler, F. and Tenenhaus-Aziza, F.<br />

D / O/ 43 A predictive model <strong>to</strong> set high pressure processing criteria 55<br />

for Listeria monocy<strong>to</strong>genes inactivation on dry-cured ham<br />

Bover-Cid, S., Belletti, N., Garriga, M. and Aymerich, T.<br />

D / O/ 44 Application <strong>of</strong> a validated predictive model <strong>to</strong> prevent growth 55<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes in ready-<strong>to</strong>-eat foods – importance<br />

for product development and risk management<br />

Mejlholm, O. and Dalgaard, P.<br />

D / O/ 45 Application <strong>of</strong> predictive microbiology <strong>to</strong> control the growth 56<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes – dairy products as an example<br />

Lobacz, A.


D / O/ 46 Characterization <strong>of</strong> Food Alert for Listeria monocy<strong>to</strong>genes in France in 2008 56<br />

Leclercq, A., Dusch, V., Salah, S., Laurent, E., Chenal-Francisque, V.,<br />

Thierry-Bled, F., Lecuit, M., Goulet, V. and Pihier, N.<br />

POSTER PRESENTATIONS<br />

D / P/ 126 Colonization <strong>of</strong> a newly constructed commercial chicken further 128<br />

processing plant with Listeria monocy<strong>to</strong>genes<br />

Berrang, M. E., Meinersmann, R., Frank, J. and Ladely, S.<br />

D / P/ 127 Episcopic differential interference contrast/epifluorescence microscopy 128<br />

<strong>to</strong> characterise in situ Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms<br />

on stainless steel surfaces<br />

Gião, M. S. and Keevil, C. W.<br />

D / P/ 128 Temperature dependent defect in bi<strong>of</strong>ilm formation 129<br />

by Listeria monocy<strong>to</strong>genes<br />

Abdalla, S., Glenn, S., Shama, G. and Andrew, P.<br />

D / P/ 129 Mode <strong>of</strong> action <strong>of</strong> Lac<strong>to</strong>coccus lactis sa31 antimicrobial peptide 129<br />

on Listeria monocy<strong>to</strong>genes ½ c<br />

Barile, M., Mormile, A., Ceres, C., Pepe, O., Cortesi, M. L. and Murru, N.<br />

D / P/ 130 Tracing the source, epidemiology and persistence <strong>of</strong> Listeria monocy<strong>to</strong>genes 130<br />

in Irish Farmhouse cheese processing facilities<br />

Jordan, K., Fox, E., O’Brien, M. and Hunt, K.<br />

D / P/ 131 Phenotypic and genotypic characteristics <strong>of</strong> Listeria monocy<strong>to</strong>genes 130<br />

strains isolated from a convenience food-processing plant<br />

Blatter, S., Stephan, R., Tasara, T. and Zweifel, C.<br />

D / P/ 132 Influence <strong>of</strong> flow direction on the adhesion <strong>of</strong> Listeria monocy<strong>to</strong>genes 131<br />

<strong>to</strong> brushed stainless steel surfaces<br />

Skovager, A., Whitehead, K., Ingmer, H., Verran, J. and Arneborg, N.<br />

D / P/ 133 Occurrence <strong>of</strong> Listeria monocy<strong>to</strong>genes in raw milk and dairy products 131<br />

in Kazerun, Iran<br />

Mehdi Mahmoodi, S. M. and Javanmardi, F.<br />

D / P/ 134 Antilisterial mode <strong>of</strong> action <strong>of</strong> bacteriocin ST182Gu produced 132<br />

by Enterococcus casseliflavus isolated from guava<br />

Todorov, S., Destro, M. T., Chiarini, E. B., Vaz-Velho, M. and Franco, B. D. G. M.<br />

D / P/ 135 Control <strong>of</strong> Listeria monocy<strong>to</strong>genes in fresh goat cheese by bacteriocinogenic 132<br />

strain Lac<strong>to</strong>coccus lactis subsp. lactis DF4Mi or commercial nisin<br />

Nader Furtado, D., Todorov, S., Landgraf, M., Destro, M. T. and Franco, B. D. G. M.<br />

D / P/ 136 High pressure processing ensures elimination <strong>of</strong> Listeria monocy<strong>to</strong>genes 133<br />

in sliced dry cured ham<br />

S<strong>to</strong>llewerk, K., J<strong>of</strong>ré, A., Comaposada, J., Aymerich, T., Ferrini, G., Arnau, J. and Garriga, M.<br />

D / P/ 137 Inhibition <strong>of</strong> Listeria monocy<strong>to</strong>genes by Lac<strong>to</strong>coccus sp. EU2241 133<br />

in tropical shrimp<br />

Abdoulaye Fall, P.<br />

D / P/ 138 Listeria monocy<strong>to</strong>genes and Listeria innocua in slaughter line <strong>of</strong> 134<br />

a swine meatpacking plant in Rio Grande do Sul State, Brazil<br />

Schittler, L. and Padilha Silva, W.<br />

D / P/ 139 Listeria monocy<strong>to</strong>genes in raw meat products marketed in the city 134<br />

<strong>of</strong> Sao Paulo, Brazil: Incidence and counts data for risk assessment<br />

Ris<strong>to</strong>ri, C. A., Rowlands, R. E. G., Martins, C. G., Fávero, L. M. and Franco, B. D. G. M.<br />

D / P/ 140 Inhibi<strong>to</strong>n <strong>of</strong> Listeria monocy<strong>to</strong>genes by Carnobacterium maltaromaticum 135<br />

in combination with extract <strong>of</strong> Lippia sidoides Cham.<br />

in cold-smoked surubim fish broth<br />

Barbosa dos Reis, F., de Souza, V. M., Sousa Thomaz, M. R., Pin<strong>to</strong> Fernandes, L.,<br />

Pereira de Oliveira, W. and Pereira De Martinis, E. C.<br />

D / P/ 141 Inhibition <strong>of</strong> Listeria monocy<strong>to</strong>genes in cooked ham 135<br />

by virulent bacteriophages and protective cultures<br />

Holck, A., Schirmer, B. C. and Berg, J.<br />

D / P/ 142 Surface colonization by Listeria monocy<strong>to</strong>genes: role <strong>of</strong> the flagella 136<br />

in the bi<strong>of</strong>ilm formation process<br />

Desvaux, M., Briandet, R., Renier, S., Deschamps, J., NCaccia, N., Chafsey, I. and Hébraud, M.<br />

D / P/ 143 The role <strong>of</strong> sanitizers in controlling Listeria monocy<strong>to</strong>genes 136<br />

on stainless steel surfaces: Lessons learned<br />

from the 2008 Listeriosis outbreak<br />

Hébert, K., Farber, J. and Pagot<strong>to</strong>, F.<br />

D / P/ 144 Listeriophage ecology and diversity on dairy farms 137<br />

Vongkamjan, K., Moreno Switt, A., den Bakker, H. C., Fortes, E. D., and Wiedmann, M.<br />

D / P/ 145 Evaluation <strong>of</strong> curative and preventive decontamination treatments 137<br />

on Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms with a new screening system<br />

Quinon, E., Chamot, S., Groelly, J., Chavant, P., Bernardi, T., Desvaux, M. and Hebraud, M.


D / P/ 146 Pulsed Field Gel Electrophoresis, conventional and molecular 138<br />

serotyping on Listeria monocy<strong>to</strong>genes: An European Pr<strong>of</strong>iciency<br />

Testing Inter-labora<strong>to</strong>ry Trial<br />

Félix, B., Roussel, S., Tam Dao, T., Asséré, A., Lombard, B. and Brisabois, A.<br />

D / P/ 147 Detection <strong>of</strong> Listeria spp. in raw and pasteurized liquid egg-products 138<br />

and in the egg-breaking plants environment<br />

Rivoal, K., Fablet, A., Chemaly, M., Salvat, G. and Protais, J.<br />

D / P/ 148 Bi<strong>of</strong>ilm formation by Listeria monocy<strong>to</strong>genes isolates under conditions 139<br />

that mimic food and digestive tract<br />

Kretli Winkelströter, L., Oliveira, M. A. and De Martinis, E. C.<br />

D / P/ 149 Bi<strong>of</strong>ilm formation <strong>of</strong> Listeria monocy<strong>to</strong>genes EGDe depends 139<br />

on temperature and nutrient availability<br />

Auchter, M., Endres, J., Waidmann, M. S. and Riedel, C. U.<br />

D / P/ 150 Antimicrobial activity <strong>of</strong> lac<strong>to</strong>coccal and enterococcal strains 140<br />

isolated from artisanal products from North West <strong>of</strong> Italy<br />

<strong>to</strong>ward Listeria monocy<strong>to</strong>genes<br />

Dal Bello, B., Rantsiou, K., Ambrosoli, R., Zeppa, G. and Cocolin, L.<br />

D / P/ 151 Plant-based strategies for Listeria monocy<strong>to</strong>genes control in foods 140<br />

Paparella, A., Serio, A., Chaves-Lopez, C. and Di Pasquale, F.<br />

D / P/ 152 Scientific studies for survival <strong>of</strong> Listeria monocy<strong>to</strong>genes 141<br />

in dairy products and its application in practice<br />

Cabanova, L., Škun<strong>to</strong>va, O. and Kantikova, M.<br />

D / P/ 153 The effect <strong>of</strong> chilling temperatures on the virulence 141<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes isolates with different origins<br />

Neves, E. M., Silva, A. C., Louro, P., Ferreira-Dias, S. and Bri<strong>to</strong>, L.<br />

D / P/ 154 How <strong>to</strong> improve a sampling plan in order <strong>to</strong> better assess 142<br />

L. monocy<strong>to</strong>genes contamination on diced bacon at the plant<br />

Bergis, H., Commeau, N., Zuliani, V., Cornu, M., Beaufort, A. and Garry, P.<br />

D / P/ 155 Contamination <strong>of</strong> Listeria monocy<strong>to</strong>genes in a cold-smoked pork 142<br />

processing plant using brining injections<br />

Berzins, A., Silins, I. and Korkeala, H.<br />

D / P/ 156 Effects <strong>of</strong> GRAS products on growth <strong>of</strong> Listeria monocy<strong>to</strong>genes 143<br />

during cold s<strong>to</strong>rage <strong>of</strong> salmon fillets<br />

McCarthy, S. and Johnson, D.<br />

D / P/ 157 Heavy-metal and detergent resistance <strong>of</strong> Listeria species isolates 143<br />

from milk processing environments<br />

Doijad, S., Garg, S. and Barbuddhe, S. B.<br />

D / P/ 158 Detection <strong>of</strong> Listeria monocy<strong>to</strong>genes in lettuce sold at markets 144<br />

and supermarkets in Por<strong>to</strong>, Portugal<br />

Noronha, L., Silva, J. and Teixeira, P.<br />

D / P/ 159 Preliminary analysis <strong>of</strong> structure and chemical composition <strong>of</strong> 144<br />

extracellular polymeric substance produced by Listeria monocy<strong>to</strong>genes<br />

Nwaiwu, O., Lad, M., Davis, A., Foster, T. and Rees, C.<br />

D / P/ 160 Ecology and persistence <strong>of</strong> Listeria monocy<strong>to</strong>genes strains in fermented 145<br />

meat sausage processors from the Northern region <strong>of</strong> Portugal<br />

Ferreira, V., Barbosa, J., Vongkamjan, K., Moreno Switt, A., Hogg, T., Gibbs, P.,<br />

Wiedmann, M. and Teixeira, P.<br />

D / P/ 161 Bi<strong>of</strong>ilm formation and survival <strong>of</strong> L. monocy<strong>to</strong>genes and slaughter house 145<br />

bacteria on surfaces at relevant environmental conditions<br />

Langsrud, S., Møretrø, T. and Heir, E.<br />

D / P/ 162 Control <strong>of</strong> L. monocy<strong>to</strong>genes by lysozyme combined with olive leaf extract 146<br />

in edible pullulan film coated on chicken breast fillets<br />

Handan Baysal, A.<br />

D / P/ 163 Evaluation <strong>of</strong> antilisterial activity by lactic acid bacteria 146<br />

Borges, S., Barbosa, J., Albano, H., Silva, J. and Teixeira, P.<br />

D / P/ 164 Molecular methods <strong>to</strong> assess Listeria monocy<strong>to</strong>genes route 147<br />

<strong>of</strong> contamination in a dairy processing plant<br />

Cocolin, L., Alessandria, V., Dolci, P. and Rantsiou, K.<br />

D / P/ 165 Persistence <strong>of</strong> L. monocy<strong>to</strong>genes in artisanal cheese producing plants 147<br />

Almeida, G., San<strong>to</strong>s, I., Magalhães, R., Barbosa, J., Hogg, T. and Teixeira, P.<br />

D / P/ 166 Occurrence <strong>of</strong> Listeria monocy<strong>to</strong>genes in food products collected 148<br />

in Portugal from retail establishments and food plants<br />

Mena, C., Carneiro, L., San<strong>to</strong>s, I., Magalhães, R., Almeida, G. and Teixeira, P.<br />

D / P/ 167 Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms grown at 12 °C showed reduced 148<br />

susceptibility <strong>to</strong> sanitizers<br />

Afonso Lourenço, A., Machado, H. and Bri<strong>to</strong>, L.<br />

D / P/ 168 Modelling growth <strong>of</strong> Listeria monocy<strong>to</strong>genes in cheese as function 149<br />

<strong>of</strong> environmental variables<br />

Sand Rosshaug, P. and Hallberg Larsen, M.<br />

D / P/ 169 Characterization <strong>of</strong> anti-Listerial bacteriocin produced by 149<br />

Lac<strong>to</strong>bacillus plantarum ST8SH, a strain isolated from Bulgarian salami<br />

Todorov, S. D. and Lemos Vaz-Velho, M.


D / P/ 170 EFSA’s proposal for an EU-wide retail survey on Listeria monocy<strong>to</strong>genes 150<br />

in selected categories <strong>of</strong> ready-<strong>to</strong>-eat food products<br />

Frank Helena Boelaert, F. H., Felício, T. and Makela, P.<br />

D / P/ 171 Ripening conditions: an asset <strong>to</strong> control L. monocy<strong>to</strong>gnes in cheeses 150<br />

Callon, C., Picque, D., Corrieu, G. and Montel, M.-C.<br />

D / P/ 172 Deterministic and s<strong>to</strong>chastic behavior <strong>of</strong> Listeria monocy<strong>to</strong>genes 151<br />

suspended cells or detached from stainless steel surfaces<br />

during cheese manufacturing<br />

Belessi, C.-E. A., Gounadaki, A. S., Arapakh, S., Schvartzman, S., Jordan, K. and Skandamis, P. N.<br />

D / P/ 173 Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in game meat 151<br />

Atanassova, V.<br />

D / P/ 174 Relationship between pathogenic pr<strong>of</strong>ile and in vitro bi<strong>of</strong>ilm formation 152<br />

capacity <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated from meat,<br />

fish and processing plants<br />

Meloni, D., Mazza, R., Marceddu, M., Piras, F., Mureddu, A. and Mazzette, R.<br />

D / P/ 175 Prevalence and molecular characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes 152<br />

in traditional fermented pork sausages produced in Italy<br />

Mazzette, R., Meloni, D., Busia, G., Melillo, R., Mureddu, A. and Piras, F.<br />

D / P/ 176 Minimum Bi<strong>of</strong>ilm Eradication Concentration (MBEC) <strong>of</strong> different 153<br />

antimicrobials on Listeria monocy<strong>to</strong>genes and Salmonella enterica bi<strong>of</strong>ilms<br />

Rodrigues, D., Teixeira, P., Oliveira, R., Ceri, H. and Azeredo, J.<br />

D / P/ 177 Examination <strong>of</strong> the ability <strong>of</strong> ad<strong>here</strong>nce, bi<strong>of</strong>ilm formation and sensitivity 153<br />

<strong>to</strong> some disinfectants <strong>of</strong> different Listeria monocy<strong>to</strong>genes strains<br />

Milanov, D., Vidić, B., Petrović, J., Bugarski, D. and Ašanin, R.<br />

D / P/ 187 Evolution <strong>of</strong> Listeria monocy<strong>to</strong>genes contamination in poultry production: 154<br />

from the farms <strong>to</strong> the processing levels<br />

Mansour Chemaly, M., Toquin, M.-T., Courtillon, C., Le Nôtre, Y., Rivoal, K. and Fravalo, P.<br />

D / P/ 188 A regular survey <strong>of</strong> Listeria in ready-<strong>to</strong>-eat foods (2004 – 2009) 154<br />

Furtado, R., Lore<strong>to</strong> Campos, M., Correia, C., Ferreira, I., Maia, C., Rosa, N., San<strong>to</strong>s, S.,<br />

San<strong>to</strong>s, M. I. and Saraiva, M.<br />

D / P/ 189 Incidence <strong>of</strong> Listeria monocy<strong>to</strong>genes in Queijo Fresco 155<br />

Rosa, N., Campos, L., Correia, C., Ferreira, I., Furtado, R., Maia, C., San<strong>to</strong>s, S.,<br />

Cunha, C. I. and San<strong>to</strong>s, M. I.<br />

D / P/ 190 Risk fac<strong>to</strong>rs for Listeria monocy<strong>to</strong>genes contamination 155<br />

in French broiler flocks<br />

Aury, K., Le Bouquin, S., Toquin, M.-T., Petetin, I., Le Nôtre, Y., Allain, V.,<br />

Fravalo, P. and Mansour Chemaly, M.<br />

D / P/ 191 Portuguese sushi: is it contaminated with L. monocy<strong>to</strong>genes? 156<br />

Mendes, D., Furtado, R., Maia, C., Correia, C., Campos Cunha, I., Pedroso, L. and San<strong>to</strong>s, M. I.<br />

D / P/ 192 Effect <strong>of</strong> the inoculum size on growth <strong>of</strong> L. monocy<strong>to</strong>genes 156<br />

in dices <strong>of</strong> poultry breast<br />

Lardeux, A.-L., Gnanou-Besse, N., Doux, C. and de Courseulles, E.<br />

D / P/ 193 Investigation in<strong>to</strong> the mechanisms <strong>of</strong> detergent induced changes 157<br />

in disinfectant susceptibility <strong>of</strong> attached Listeria monocy<strong>to</strong>genes<br />

Wal<strong>to</strong>n, J., Hayes, R., Protheroe, R., Hill, D. and Gibson, H.<br />

D / P/ 194 Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes throughout the production process 157<br />

<strong>of</strong> Parma ham: tracing contaminations from slaughterhouses<br />

<strong>to</strong> the final product<br />

Prencipe, V. A., Rizzi, V., Iannetti, L., Serraino, A., Calderone, D., Rossi, A., Morelli, D.,<br />

Marino, L. and Migliorati, G.<br />

D / P/ 195 Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in raw milk sold at vending machines 158<br />

in Abruzzo region<br />

Prencipe, V. A., Scat<strong>to</strong>lini, S., Sperandii, A. F. and Migliorati, G.<br />

D / P/ 196 Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated from s<strong>of</strong>t 158<br />

and semi s<strong>of</strong>t cheeses sampled at retail level<br />

Acciari, V., Torresi, M., Migliorati, G., Di Giannatale, E., Semprini, P. and Prencipe, V.<br />

D / P/ 197 Preliminary report on the organisation <strong>of</strong> a food microbiology 159<br />

pr<strong>of</strong>iciency testing program as a <strong>to</strong>ol <strong>to</strong> guarantee the equivalence<br />

<strong>of</strong> the US and IT <strong>of</strong>ficial control systems<br />

Di Giannatale, E., Marfoglia, C., Prencipe, V., Salini, R., Migliorati, G. and Ricci, L.<br />

D / P/ 198 High nisin susceptibility <strong>of</strong> Listeria spp. wild-type strains isolated 159<br />

from dairies with traditional cheese preservation in Portugal<br />

Pintado, C. M. B. S and Ferreira M. A. S. S.<br />

D / P/ 199 Utilization <strong>of</strong> Lac<strong>to</strong>coccus lactis M104, a wild nisin-producing 163<br />

raw milk isolate, as an antilisterial adjunct in traditional<br />

Greek Graviera cheese processing<br />

Samelis, I., Pappa, E., Bogovic-Matijasic, B. and Rogelj, I.<br />

D / P/ 200 The European Project BASELINE “Selection and improving 163<br />

<strong>of</strong> fit-for-purpose sampling procedures for specific foods and risks”<br />

Manfreda, G. and De Cesare, A.


AREA //E // Communication, risk perception and consumer practices –<br />

Social sciences in Listeria control<br />

PLENARY LECTURES<br />

E / PL/ 13 How can the social sciences help us understand the prevalence <strong>of</strong> 29<br />

listeriosis in the UK?<br />

Wadge, A.<br />

E / PL/ 14 Consumer perceptions, behaviour and microbial food safety; 29<br />

implications for Listeria control<br />

Frewer, L. J.<br />

ORAL PRESENTATIONS<br />

E / O/ 47 Efforts <strong>to</strong> update and perform health education on listeriosis in 57<br />

Los Angeles County, California, by the County <strong>of</strong> Los Angeles Department<br />

<strong>of</strong> Public Health<br />

Guevara, R. E.<br />

E / O/ 48 Awareness <strong>of</strong> listeriosis among Portuguese pregnant women 57<br />

Mateus, T., Maia, R. L. and Teixeira, P.<br />

E / O/ 49 The development and progress <strong>of</strong> a risk-based strategy 58<br />

for the control <strong>of</strong> Listeria monocy<strong>to</strong>genes in New Zealand<br />

Castle, M. and Crerar, S.<br />

POSTER PRESENTATIONS<br />

E / P/ 178 What is an appropriate level <strong>of</strong> protection for Listeria monocy<strong>to</strong>genes 160<br />

in foodstuffs consumed by vulnerable groups?<br />

Little, C., Gillespie, I., Grant, K., Gormley, F., Mook, P. and McLauchlin, J.<br />

E / P/ 179 ILCD: An Interactive Listeria culture diversity knowledgebase 160<br />

Ashok Kumar, J., Barbuddhe, S. B., Kalekar, S., Rodrigues, J., Chopade, N. A.,<br />

Hain, T. and Chakraborty, T.<br />

E / P/ 180 Listeria spp. and the domestic environment: consumer knowledge, 161<br />

attitudes, risk perceptions and food-handling behaviours<br />

Redmond, E. C.<br />

E / P/ 181 Do you know the temperature in your refrigera<strong>to</strong>r? 161<br />

Røssvoll, E., Jacobsen, E., Ueland, Ø., Einar Granum, P. and Langsrud, S.<br />

Authors Index 165


PLENARY LECTURES


KEY NOTE SPEACH<br />

Listeria monocy<strong>to</strong>genes: a multifaceted model<br />

Cossart, P.*<br />

Institut Pasteur, Unité des Interactions Bactéries cellules, Inserm U604, INRA USC2020<br />

Listeria monocy<strong>to</strong>genes is an intracellular pathogen responsible for severe human<br />

food-borne infections characterized by gastroenteritis, materno-fetal infections<br />

and brain infections with a mortality rate <strong>of</strong> 30 %. The disease is mainly due the capacity<br />

<strong>of</strong> Listeria <strong>to</strong> cross three host barriers: the intestinal barrier, the matern<strong>of</strong>etal<br />

barrier and the blood brain barrier. It is also due <strong>to</strong> the capacity <strong>of</strong> the organism<br />

<strong>to</strong> survive and replicate in macrophages and <strong>to</strong> enter and replicate in non<br />

phagocytic cells. We use a combination <strong>of</strong> approaches <strong>to</strong> understand the mechanisms<br />

which allow establishment and maintenance <strong>of</strong> a Listeria infection. We then<br />

investigate if our findings have a general significance. In nearly three decades <strong>of</strong><br />

molecular and cellular investigations, the study <strong>of</strong> Listeria momocy<strong>to</strong>genes and <strong>of</strong><br />

listeriosis has led <strong>to</strong> new concepts in infection biology and in several other areas <strong>of</strong><br />

biology including cell biology, microbiology, molecular medicine and genomics.<br />

* Plenary Supported by FCT<br />

21


R E F E R E N C E<br />

A / PL<br />

01<br />

R E F E R E N C E<br />

A / PL<br />

02<br />

22<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

PLENARY LECTURES // PL/ 1–2<br />

The pangenome <strong>of</strong> Listera spp.<br />

Chakraborty, T.*<br />

Institute <strong>of</strong> Medical Microbiology, Justus-Liebig University <strong>of</strong> Giessen, Germany<br />

Listeria monocy<strong>to</strong>genes is a food-borne pathogen with a high mortality rate that<br />

has served as an invaluable model for intracellular parasitism. We have sequenced<br />

representative genomes comprising all species for the genus Listeria as well as<br />

strains representing clonal lineages <strong>of</strong> the pathogenic species <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

Comparative genome analysis now provides clear evidence indicating that the various<br />

non-pathogenic species <strong>of</strong> Listeria have been derived by gene loss and/or mutational<br />

decay <strong>of</strong> virulence- and niche-adaptive fac<strong>to</strong>rs from a progeni<strong>to</strong>r strain<br />

that harboured many <strong>of</strong> the currently known virulence fac<strong>to</strong>rs. Thus, non-pathogenic<br />

Listeria are compromised with regard <strong>to</strong> their ability <strong>to</strong> live in the cy<strong>to</strong>plasm<br />

<strong>of</strong> infected host cells but have acquired genes enabling their growth in soil and decaying<br />

vegetation. Comparative transcrip<strong>to</strong>me analysis <strong>of</strong> intracellular growth has<br />

also uncovered additional levels <strong>of</strong> adaptive evolution in growth among the different<br />

lineages <strong>of</strong> L. monocy<strong>to</strong>genes. Analysis <strong>of</strong> the pan-genome <strong>of</strong> L. monocy<strong>to</strong>genes<br />

strains revealed an extensive, as yet unexplained gene-reper<strong>to</strong>ire in these<br />

genomes, and provides evidence for the evolution <strong>of</strong> these strains by gathering<br />

genes from organisms in different environmental and host niches.<br />

* Plenary Supported by FCT<br />

Ecology <strong>of</strong> L. monocy<strong>to</strong>genes and Listeria spp. in natural<br />

and food associated environments<br />

Wiedmann, M.*<br />

Department <strong>of</strong> Food Science, Cornell University, Ithaca, NY, USA<br />

Listeria spp., including the pathogen L. monocy<strong>to</strong>genes, can be isolated from a variety<br />

<strong>of</strong> environments, <strong>of</strong>ten at considerable frequency. A number <strong>of</strong> our studies<br />

have specifically shown that Listeria spp. can <strong>of</strong>ten be isolated from > 20 % <strong>of</strong> samples<br />

collected from natural, urban, and farm environments and can also be commonly<br />

isolates from food associated environments (e.g., processing plants, retail<br />

environments). Molecular characterization and subtyping <strong>to</strong>ols not only typically<br />

reveal considerable diversity among Listeria spp. isolates from different environments,<br />

but also provide evidence for (i) association between certain subtypes or<br />

species and specific environments (suggesting existence <strong>of</strong> specific Listeria ecotypes)<br />

as well as (ii) long-term persistence <strong>of</strong> specific Listeria strains in different<br />

environments. For example, in one <strong>of</strong> our studies, L. seeligeri and L. welshimeri<br />

were significantly associated with natural environments (p


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

PLENARY LECTURES // PL/ 3–6<br />

How Listeria monocy<strong>to</strong>genes breaches host barriers<br />

Lecuit, M.*<br />

Institut Pasteur, Inserm, Paris, France<br />

Listeria monocy<strong>to</strong>genes (Lm) is a human foodborne pathogen that causes listeriosis,<br />

a systemic infection leading <strong>to</strong> meningitis, encephalitis and fe<strong>to</strong>-placental infection.<br />

To reach its target organs, Lm crosses the intestinal, blood-brain and placental<br />

barriers. We will present the results <strong>of</strong> our investigations regarding the<br />

molecular mechanisms underlying Lm crossing <strong>of</strong> these three barriers.<br />

* Plenary Supported by FCT<br />

Secretion <strong>of</strong> a novel L. monocy<strong>to</strong>genes cyclic dinucleotide in<strong>to</strong> the<br />

cy<strong>to</strong>sol <strong>of</strong> infected host cells activates an innate immune pathway<br />

Portnoy, D. A.*<br />

Department <strong>of</strong> Molecular and Cell Biology and The School <strong>of</strong> Public Health, University <strong>of</strong> California,<br />

Berkeley, USA<br />

Listeria monocy<strong>to</strong>genes has been used for decades as a model <strong>to</strong> study basic aspects<br />

<strong>of</strong> intracellular parasitism and cell-mediated immunity. Because L. monocy<strong>to</strong>genes<br />

induces a robust CD8+ T-cell response, attenuated strains <strong>of</strong> L. monocy<strong>to</strong>genes<br />

are being developed as live, attenuated vaccine vec<strong>to</strong>rs for infectious<br />

disease and malignancies. What makes L. monocy<strong>to</strong>genes such a strong inducer <strong>of</strong><br />

cellular immunity? As everyone in this audience appreciates, virulent strains <strong>of</strong><br />

L. monocy<strong>to</strong>genes secrete a pore-forming cy<strong>to</strong>lysin (LLO) that allows bacteria access<br />

the host cell cy<strong>to</strong>sol. We previously discovered that vacuolar and cy<strong>to</strong>solic<br />

bacteria induced distinct host transcriptional responses, the latter leading <strong>to</strong> the<br />

expression <strong>of</strong> beta interferon and a host <strong>of</strong> co-regulated genes. To understand the<br />

microbial components that activate the cy<strong>to</strong>solic pathway, we used a forward genetic<br />

screen <strong>to</strong> identify bacterial mutants that induced an enhanced or diminished<br />

host transcriptional response <strong>to</strong> cy<strong>to</strong>solic bacteria. Most <strong>of</strong> the mutants identified<br />

mapped <strong>to</strong> genes encoding or regulating multidrug efflux pumps. We now have<br />

a series <strong>of</strong> strains that induce beta interferon over a 60-fold range. These data<br />

were consistent with a model in which a small molecula is either actively or inadvertently<br />

being pumped from cells, and that the host can recognize and respond.<br />

Using conventional biochemistry and mass spectrometry, we have identified the<br />

L. monocy<strong>to</strong>genes molecule as a cyclic dinucleotide. The precise structure <strong>of</strong> this<br />

novel bacterial molecule, identification <strong>of</strong> the cyclase and phosphodiesterase will<br />

be presented at the meeting.<br />

* Plenary Supported by FLAD<br />

R E F E R E N C E<br />

B / PL<br />

03<br />

R E F E R E N C E<br />

B / PL<br />

04<br />

23


R E F E R E N C E<br />

B / PL<br />

05<br />

R E F E R E N C E<br />

B / PL<br />

06<br />

24<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

PLENARY LECTURES // PL/ 3–6<br />

Diagnosis and clinical management <strong>of</strong> listeriosis<br />

in ruminants and camelids<br />

Poulsen, K. P.*<br />

School <strong>of</strong> Veterinary Medicine, University <strong>of</strong> Wisconsin, Madison, USA<br />

Listeria monocy<strong>to</strong>genes is a successful intracellular pathogen <strong>of</strong> domestic and food<br />

producing animals. Ruminant (cattle, sheep, goats) and camelid (llama and alpaca)<br />

species are constantly exposed <strong>to</strong> this gram-positive pathogen. It is ubiqui<strong>to</strong>us in<br />

the environment and tends <strong>to</strong> overgrow in poorly prepared silage (pH > 5.0). Listeriosis<br />

manifests in these species as three distinct clinical syndromes including<br />

abortion, neonatal sepsis, and meningioencephalitis (circling disease). Infection in<br />

food producing animals, and subsequent shedding <strong>of</strong> L. monocy<strong>to</strong>genes in milk and<br />

meat represents a significant risk <strong>to</strong> food safety. In this review, clinical signs, diagnostics,<br />

and treatment <strong>of</strong> listeriosis <strong>of</strong> ruminant and camelid species will be presented<br />

including video clips <strong>of</strong> clinically affected animals.<br />

* Plenary Supported by FLAD<br />

Listeriolysin S – a second haemolysin with a role in the virulence<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Hill, C.*<br />

Alimentary Pharmabiotic Centre and Microbiology Department, University College Cork, Ireland<br />

Listeria monocy<strong>to</strong>genes require listeriolysin O (encoded by llo or hlyA) <strong>to</strong> escape<br />

the vacuole and initiate intracellular growth and intercellular spread. Llo - mutants<br />

are essentially avirulent and so listeriolysin O is regarded as the primary virulence<br />

fac<strong>to</strong>r in L. monocy<strong>to</strong>genes. Given that all strains and serotypes possess this<br />

virulence fac<strong>to</strong>r, it is unlikely <strong>to</strong> explain the increased incidence <strong>of</strong> particular<br />

serotypes (such as 4b) in listeriosis epidemics. Many labora<strong>to</strong>ries have sought <strong>to</strong><br />

solve the Listeria conundrum <strong>of</strong> why certain strains may be more likely <strong>to</strong> cause<br />

disease in humans. We have identified a second haemolysin, which we have named<br />

listeriolysin S, which is associated with epidemic strains. Listeriolysin S is very<br />

different from listeriolysin O in that it is a highly modified peptide, which resembles<br />

(in probable structure if not in primary sequence) the strep<strong>to</strong>coccal virulence<br />

fac<strong>to</strong>r, strep<strong>to</strong>lysin S. In fact, a family <strong>of</strong> these modified virulence peptides can be<br />

inferred from genomic data <strong>of</strong> other pathogens, including Staphylococcus aureus<br />

and Clostridium botulinum. listeriolysin S is not expressed under labora<strong>to</strong>ry conditions,<br />

but can be induced with a number <strong>of</strong> reagents in vitro. In addition, a mutant<br />

in which listeriolysin S is constitutively expressed is hyper-hemolytic in comparison<br />

<strong>to</strong> wildtype strains. We have also demonstrated a small, but significant,<br />

role for listeriolysin O in murine models <strong>of</strong> infection and in polymorphonuclear<br />

leucocyte survival assays. Listeriolysin S provides a possible explanation for the<br />

enhanced virulence <strong>of</strong> certain strains in human listeriosis, but much remains <strong>to</strong> be<br />

done <strong>to</strong> confirm or refute this hypothesis.<br />

* Plenary Supported by FEMS


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

PLENARY LECTURES // PL/ 7–9<br />

Update on clinical aspects <strong>of</strong> listeriosis: What’s new?<br />

Schlech, W. F.*<br />

Dalhousie University, Canada<br />

Listeriosis remains a challenging problem for clinicians. Diagnosis may be delayed<br />

and treatments prolonged although decreases in mortality have been noted in<br />

some surveillance programs. New syndromes, such as necrotizing fasciitis, have<br />

been described as well as further outbreaks <strong>of</strong> febrile gastroenteritis. Sepsis and<br />

rhomboencephalitis remain the most common presenting syndromes for invasive<br />

listeriosis. Genetic markers for infection have also been uncovered. Infection in<br />

the immunocompetent host still occurs but risk fac<strong>to</strong>rs for the disease primarily<br />

points <strong>to</strong> abnormalities in cell-mediated and innate immunity as major predispositions<br />

<strong>to</strong> listeriosis. Use <strong>of</strong> TNF-alpha inhibi<strong>to</strong>rs and corticosteroids are <strong>of</strong> particular<br />

concern. Some newer antibiotics active against L. monocy<strong>to</strong>genes have been<br />

studied but ampicillin and gentamicin remains the treatment <strong>of</strong> choice. T<strong>here</strong> is<br />

further evidence that trimethoprim-sulfamethoxazole has a strong protective effect<br />

against listeriosis in the compromised host receiving this drug for protection<br />

against other pathogens.<br />

* Plenary Supported by FCT<br />

A Canadian outbreak <strong>of</strong> listeriosis due <strong>to</strong> deli-meat:<br />

Driving change in the food safety system<br />

Farber, J. M. 1*, Pagot<strong>to</strong>, F. 1, Gilmour, M. 2, Nadon, C. 2, Savelli, C. 3, and MacDonald, D. 3<br />

1. Food Direc<strong>to</strong>rate, Health Canada<br />

2. National Microbiology Labora<strong>to</strong>ry, Public Health Agency Canada<br />

3. Centre for Food-borne, Environmental and Zoonotic Infectious Diseases,<br />

Public Health Agency Canada<br />

In 2008, Canada experienced the first multi-provincial and largest outbreak <strong>of</strong> invasive<br />

listeriosis. During this outbreak, 57 cases <strong>of</strong> illness, in 7 provinces, resulted<br />

in 23 deaths from listeriosis, which was traced back <strong>to</strong> contaminated deli-meat<br />

from Company A. The age range was 29 <strong>to</strong> 98 years, with the median age being 78.<br />

All <strong>of</strong> the cases with a known medical his<strong>to</strong>ry prior <strong>to</strong> exposure had underlying<br />

medical conditions. Fifty (88 %) cases reported deli-meat consumption. Additionally,<br />

84 % <strong>of</strong> cases had institutional exposure. Investiga<strong>to</strong>rs confirmed Company A<br />

deli meat was served <strong>to</strong> 27 cases. The major probable cause <strong>of</strong> the Listeria contamination<br />

in Company A’s plant was related <strong>to</strong> meat slicing equipment. High-throughput<br />

genome sequencing <strong>of</strong> two serotype 1/2a Listeria monocy<strong>to</strong>genes isolates was<br />

completed during the outbreak. By screening for genetic traits specific <strong>to</strong> the outbreak<br />

genomes, examination <strong>of</strong> clinical, environmental and food isolates associated<br />

with the outbreak revealed that three distinct, but highly-related strains may have<br />

been involved in this nationwide outbreak. As a result <strong>of</strong> the outbreak, the Public<br />

Health Agency <strong>of</strong> Canada (PHAC), the Canadian Food Inspection Agency (CFIA)<br />

and Health Canada (HC) conducted independent lessons-learned exercises. In addition,<br />

the CFIA developed manda<strong>to</strong>ry new Directives for federally-registered meat<br />

and poultry plants, and Health Canada started updating their policy on Listeria<br />

monocy<strong>to</strong>genes in RTE foods. Furthermore, a federal review by an independent investiga<strong>to</strong>r<br />

(Ms. Sheila Weatherill) in<strong>to</strong> the outbreak resulted in 57 recommendations<br />

directed at food processors, regula<strong>to</strong>rs, public health pr<strong>of</strong>essionals and individual<br />

consumers. Work is currently underway at PHAC, HC and CFIA <strong>to</strong> directly<br />

address the recommendations coming out <strong>of</strong> the Weatherill report.<br />

* Plenary Supported by FCT<br />

R E F E R E N C E<br />

C / PL<br />

07<br />

R E F E R E N C E<br />

C / PL<br />

08<br />

25


R E F E R E N C E<br />

C / PL<br />

09<br />

26<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

PLENARY LECTURES // PL/ 7–9<br />

Listeria monocy<strong>to</strong>genes phagocytic strategy<br />

Alvarez-Dominguez, C.*<br />

Hospital Santa Cruz de Liencres-IFIMAV and IES Zapa<strong>to</strong>n, Spain<br />

Listeria monocy<strong>to</strong>genes enters macrophages and resides for a short period within<br />

the phagosomal compartment before escaping and replicating in the cy<strong>to</strong>sol. Listeria<br />

has evolved a fine phagosomal strategy <strong>to</strong> survive the microbicidal machinery<br />

<strong>of</strong> macrophages and avoid eliciting a strong innate immune response. First,<br />

Listeria-GAPDH enzymatic modification and inactivation <strong>of</strong> Rab5a caused a delay<br />

on phagosome maturation. Second, Listeria interferes with the trafficking <strong>of</strong> two<br />

lysosomal proteins, cathepsin-D and LIMP2. Cathepsin-D enzymatic action blocks<br />

the pore-forming function <strong>of</strong> Listeria-LLO within the phagosomes. T<strong>here</strong>fore, inhibiting<br />

the transport <strong>of</strong> this lysosomal protease <strong>to</strong> the phagosomes increases Listeria<br />

viability. Finally, the Listeria interference with LIMP2 trafficking disrupts<br />

the connection between late endosomal events and the onset <strong>of</strong> innate immunity.<br />

In brief, Listeria phagosomal strategy has the purpose <strong>to</strong> avoid the phagosomal<br />

transformation in<strong>to</strong> fully competent bactericidal and antigen-processing compartments.<br />

Deciphering Listeria phagosomal strategy <strong>to</strong> subvert the host immune<br />

response might help <strong>to</strong> design better therapies against listeriosis.<br />

* Plenary Supported by FCT


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

PLENARY LECTURES // PL/ 10–12<br />

Risk assessment using the microbiological criteria:<br />

Arguments for zero <strong>to</strong>lerance (USDA) other viewpoint (Europe):<br />

Risk-based microbiological criteria for Listeria monocy<strong>to</strong>genes<br />

in RTE foods<br />

Luber, P.*<br />

Federal Office <strong>of</strong> Consumer Protection and Food Safety, Berlin, Germany<br />

The control <strong>of</strong> Listeria monocy<strong>to</strong>genes in foods has been a focus activity ever since<br />

the link between listeriosis in vulnerable populations and exposure via foods became<br />

clear. Listeria spp. originate from soil and can contaminate many different<br />

foods during food production or via processing environments. Once Listeria bacteria<br />

get in a food plant, they might survive in the processing environment over<br />

long time periods and could pose a contamination hazard <strong>to</strong> produced foods. Moreover,<br />

the ability <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> multiply under refrigeration temperatures<br />

and without oxygen, for example in vacuum packs or in foods which are packed in<br />

modified atmosp<strong>here</strong>s, makes it challenging <strong>to</strong> control the bacteria further up in<br />

the food chain at the retail level and once foods have reached consumers’ homes.<br />

Risk assessments done in the past years have shown that amongst the many foods<br />

which can become contaminated with L. monocy<strong>to</strong>genes, ready-<strong>to</strong>-eat (RTE) foods<br />

which are consumed without further heat treatment are associated with the greatest<br />

listeriosis risk. Regula<strong>to</strong>rs in the European Community and the Codex Alimentarius<br />

Commission have developed risk-based microbiological criteria for L. monocy<strong>to</strong>genes<br />

in foods. In both cases, the microbiological criteria specifically apply <strong>to</strong><br />

RTE foods and take in<strong>to</strong> consideration if growth <strong>of</strong> Listeria in the food may occur<br />

or not. However, in both approaches the main focus is on controlling L. monocy<strong>to</strong>genes<br />

during processing <strong>of</strong> RTE foods and in the production environment. The European<br />

Regulation on microbiological criteria (Regulation (EC) no 2073/2005) is<br />

embedded in the so called ‘hygiene package’ <strong>of</strong> legislations, and in the Codex Alimentarius<br />

microbiological criteria are presented in an Annex <strong>of</strong> the ‘Guidelines<br />

on the Application <strong>of</strong> General Principles <strong>of</strong> Food Hygiene <strong>to</strong> the Control <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes in Ready-<strong>to</strong>-Eat foods’ (CAC/GL 61-2007), only. In the European<br />

Community, food business opera<strong>to</strong>rs have an obligation <strong>to</strong> demonstrate compliance<br />

with microbiological criteria and t<strong>here</strong>by enable verification <strong>of</strong> their GMP<br />

and HACCP systems. The approach <strong>of</strong> using risk-based microbiological criteria<br />

for L. monocy<strong>to</strong>genes as management option for controlling Listeria in foods and<br />

thus <strong>to</strong> reduce the likelihood <strong>of</strong> listeriosis in vulnerable populations will be critically<br />

discussed.<br />

* Plenary Supported by FCT<br />

USDA regula<strong>to</strong>ry approach and considerations for the control<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Engeljohn, D.*<br />

U.S. Department <strong>of</strong> Agriculture, Food Safety and Inspection Service, Deputy Assistant Administra<strong>to</strong>r,<br />

Office <strong>of</strong> Policy and Program Development, USA<br />

An overview <strong>of</strong> the design <strong>of</strong> the risk management design and objectives <strong>of</strong> the<br />

control programs for ready-<strong>to</strong>-eat meat and poultry will be presented. Use <strong>of</strong> risk<br />

assessments <strong>to</strong> inform the verification and enforcement strategy will be highlighted.<br />

Lessons learned regarding the implementation <strong>of</strong> a “zero <strong>to</strong>lerance” approach<br />

will be discussed, along with future plans for ensuring that progress <strong>to</strong>wards<br />

good control are not negated in further processing operations.<br />

* Plenary Supported by FLAD<br />

R E F E R E N C E<br />

D / PL<br />

10<br />

R E F E R E N C E<br />

D / PL<br />

11<br />

27


R E F E R E N C E<br />

D / PL<br />

12<br />

28<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

PLENARY LECTURES // PL/ 10–12<br />

Listeria monocy<strong>to</strong>genes, an emergent pathogen<br />

in Chile and Latin America<br />

Hormazábal, J. C.*<br />

Public Health Institute <strong>of</strong> Chile<br />

Food borne diseases are a major health burden in Latin America. In Chile Listeria<br />

monocy<strong>to</strong>genes is a pathogen under labora<strong>to</strong>ry surveillance since 2004. All the isolates<br />

from human cases must be confirmed at the National Reference Labora<strong>to</strong>ry<br />

at the Public Health Institute <strong>of</strong> Chile. Until 2008 L. monocy<strong>to</strong>genes was an infrequent<br />

pathogen, approximately 45 clinical cases per year were reported in all the<br />

country. At the ending <strong>of</strong> 2008 an unusual increase <strong>of</strong> listeriosis was detected in<br />

the Metropolitan region <strong>of</strong> Chile. In this scenario, the Reference Labora<strong>to</strong>ry included<br />

for first time molecular <strong>to</strong>ols for the detection <strong>of</strong> related cases. PFGE including<br />

PulseNet <strong>International</strong> pro<strong>to</strong>cols, was a powerful upgrade for the surveillance<br />

system. In parallel The Ministry <strong>of</strong> health, in absence <strong>of</strong> a specific regula<strong>to</strong>ry<br />

framework for L. monocy<strong>to</strong>genes in food industry, made important changes in sanitary<br />

food regulations, including specific microbiological criteria for Listeria in<br />

food and environment. The integration <strong>of</strong> surveillance data, clinical, environmental<br />

and food industry isolates, allowed the creation <strong>of</strong> a single database, including<br />

molecular typing information. This dynamic system allowed the detection <strong>of</strong> the<br />

first Chilean outbreak <strong>of</strong> L. monocy<strong>to</strong>genes, and made possible the link <strong>of</strong> clinical<br />

cases <strong>to</strong> potential sources, including the environment, raw or processed food products.<br />

After few weeks <strong>of</strong> an intensive investigation the outbreak source was confirmed<br />

(goat and brie cheese). The inclusion <strong>of</strong> molecular <strong>to</strong>ols in surveillance<br />

provides valuable information for the early detection <strong>of</strong> listeriosis outbreaks.<br />

PulseNet Latin America Network gives an important support for the construction<br />

<strong>of</strong> national and regional databases for the detection <strong>of</strong> local outbreaks and virulent<br />

clones with potential international spread. In Latin America Listeria keeps as an<br />

infrequent and unknown pathogen, major efforts are necessary specially in health<br />

staff and general population education including a sensitive surveillance system<br />

and a strong regula<strong>to</strong>ry framework in food industry.<br />

* Plenary Supported by FCT


AREA // E // Communication, risk perception and consumer practices – Social sciences in Listeria control<br />

PLENARY LECTURES // PL/ 13–14<br />

How can the social sciences help us understand the prevalence<br />

<strong>of</strong> listeriosis in the UK?<br />

Wadge, A.*<br />

Chief Scientist, Food Standards Agency, UK<br />

T<strong>here</strong> has been a marked change in the epidemiology <strong>of</strong> listeriosis in humans in the<br />

UK over the past decade. A doubling <strong>of</strong> reported cases has been seen in England and<br />

Wales since 2001 and this has occurred largely in people aged 60 years and over and<br />

presenting with bacteraemia rather than central nervous system infection. The incidence<br />

<strong>of</strong> listeriosis among other age ranges has remained unchanged and the incidence<br />

in pregnant women has been stable since the early 1990s. This presentation<br />

will explore the possible reasons for the increase in the over 60s focussing on several<br />

strands <strong>of</strong> work that are being undertaken <strong>to</strong> address this. Firstly in 2007 the UK<br />

Advisory Committee on the Microbiological Safety <strong>of</strong> Food (ACMSF) considered<br />

that the change in the epidemiology seen in the UK was most likely linked <strong>to</strong> social<br />

fac<strong>to</strong>rs rather than changes occurring in the microorganism and referred the issue <strong>of</strong><br />

listeriosis in the elderly <strong>to</strong> its ad hoc group on vulnerable groups for further consideration.<br />

The presentation will highlight some <strong>of</strong> the group’s findings which are available<br />

in a report (www.food.gov.uk/multimedia/pdfs/committee/acmsflisteria.pdf ).<br />

Amongst their recommendations they recognised the importance <strong>of</strong> food consumption<br />

and food handling behaviours in the over 60s including those in vulnerable<br />

groups. One <strong>of</strong> the report’s recommendations was for this <strong>to</strong> be considered by the<br />

Food Standards Agency’s (FSA) Social Science Research Committee (SSRC). The<br />

presentation will highlight findings from the SSRC’s report which showed that very<br />

little is known about food s<strong>to</strong>rage and handling practices <strong>of</strong> over 60s in the home.<br />

Studies suggest that older people handle food differently from younger people although<br />

the reason for this difference is unclear. Specific differences reported included<br />

poor refrigeration and defrosting practices; differences in cooling, s<strong>to</strong>rage<br />

and reheating <strong>of</strong> lef<strong>to</strong>vers; variable ad<strong>here</strong>nce <strong>to</strong> ‘use-by’ and ‘best-before’ dates; and<br />

differences in personal and domestic hygiene such as hand-washing and cleaning <strong>of</strong><br />

kitchen surfaces. The report noted that t<strong>here</strong> is relatively little evidence regarding<br />

the current levels <strong>of</strong> food hygiene knowledge among those aged 60 years and over<br />

and no information on whether the level <strong>of</strong> knowledge differs with generations or<br />

has changed as people age. Emphasising the need for correct s<strong>to</strong>rage and handling <strong>of</strong><br />

food in the home particularly by those over 60s was a key theme <strong>of</strong> Food Safety week<br />

in the UK in 2009 and the presentation will illustrate the approach that was taken by<br />

the FSA in this campaign. The presentation will conclude by highlighting some <strong>of</strong><br />

the gaps in our knowledge and the SSRC recommendations concerning research.<br />

* Plenary Supported by FCT<br />

Consumer perceptions, behaviour and microbial food safety;<br />

implications for Listeria control<br />

Frewer, L. J.*<br />

Wageningen University, The Netherlands<br />

In order <strong>to</strong> understand fully the problem <strong>of</strong> food safety linked <strong>to</strong> the occurrence <strong>of</strong> listeriosis,<br />

it is important <strong>to</strong> consider how consumers respond <strong>to</strong> food safety issues, in terms<br />

<strong>of</strong> their psychology and how this determines their behaviour, for example in relation <strong>to</strong><br />

food preparation behaviour. An important research objective relates <strong>to</strong> consumer activities<br />

after food purchase. Risk perception determines how consumers react <strong>to</strong> different<br />

types <strong>of</strong> risks. Very generally, risks which are perceived <strong>to</strong> be unnatural in origin<br />

and involuntarily imposed on the individual who is exposed <strong>to</strong> them are perceived <strong>to</strong> be<br />

more threatening. Microbiological risks are perceived <strong>to</strong> be “naturally occurring” and,<br />

in the case <strong>of</strong> food safety risks, highly controllable, and so are not a focus <strong>of</strong> consumer<br />

concern. In addition, consumers tend <strong>to</strong> exhibit an “optimistic bias” in relation <strong>to</strong> microbiological<br />

food risks, which means that food safety information tends <strong>to</strong> be perceived<br />

as applying <strong>to</strong> more vulnerable, consumers in the population. Research suggests<br />

that consumers are reasonably knowledgeable about safe food preparation practices,<br />

but that this knowledge is not always applied in practice. Food preparation tends <strong>to</strong> be a<br />

habitual behaviour, which is difficult <strong>to</strong> change. Introducing food safety messages during<br />

food preparation (for example, in recipe development) tends <strong>to</strong> activate existing<br />

food safety knowledge, as does the inclusion <strong>of</strong> materials designed <strong>to</strong> elicit affective responses<br />

<strong>to</strong> food safety issues, for example disgust. Some groups within the population<br />

are more vulnerable than others, and targeting risk communication messages <strong>to</strong> the<br />

needs <strong>of</strong> these groups is particularly relevant. It is argued that, whilst consumer observation<br />

studies, combined with modelling <strong>of</strong> critical control points in food preparation,<br />

might indicate the riskiest behaviours in terms <strong>of</strong> Listeria, this information should be<br />

combined with activation <strong>of</strong> general food safety knowledge if an effective approach <strong>to</strong><br />

reducing the incidence <strong>of</strong> food borne disease is <strong>to</strong> be developed.<br />

* Plenary Supported by FCT<br />

R E F E R E N C E<br />

E / PL<br />

13<br />

R E F E R E N C E<br />

E / PL<br />

14<br />

29


ORAL PRESENTATIONS


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

The cold shock associated proteins (Csps) promote <strong>to</strong>lerance <strong>of</strong> different<br />

environmental stresses and host cell invasion <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Tasara, T. 1, Klumpp, J. 2, Loessner, M. J. 2 and Stephan, R. 1<br />

1. Institute <strong>of</strong> Food Safety, University <strong>of</strong> Zurich, Switzerland<br />

2. Institute <strong>of</strong> Food, Nutrition and Health, ETH Zurich, Switzerland<br />

The food-borne pathogen Listeria monocy<strong>to</strong>genes has <strong>to</strong> withstand various stress<br />

conditions in order <strong>to</strong> survive and proliferate on foods and within different types<br />

<strong>of</strong> host cells. The bacterial cold shock family proteins (Csps) consist <strong>of</strong> small,<br />

highly conserved nucleic acid-binding proteins, and are assumed <strong>to</strong> have an influence<br />

on the expression <strong>of</strong> various microbial genes. In addition <strong>to</strong> cold adaptation<br />

functions, some prokaryotic Csp proteins are also involved in promotion <strong>of</strong> other<br />

cell processes during normal bacterial growth, as well as in adaptation <strong>to</strong> nutrient<br />

starvation and stationary growth phase stresses. The possible functional contribution<br />

<strong>of</strong> L. monocy<strong>to</strong>genes CspA, CspB and CspD proteins were investigated<br />

under food-related environmental stress, and during host cell invasion. We show<br />

that the three L. monocy<strong>to</strong>genes Csp components, although highly homologous,<br />

provide distinct cellular functions during cold, osmotic and oxidative stress adaptation<br />

as well as host cell invasion processes. All three csp genes are constitutively<br />

expressed but are dispensable for viability and growth <strong>of</strong> this organism at optimal<br />

temperature. However, a hierarchy in Csp functional importance during cold<br />

(CspA > CspD > CspB) and osmotic (CspD > CspA / CspB) stress adaptation is observed.<br />

With respect <strong>to</strong> double (DcspBD) and triple (DcspABD) csp deletion mutants,<br />

we also observe reduced survival <strong>of</strong> oxidative stress and a reduced ability <strong>to</strong><br />

invade Caco-2 and J744A.1 murine macrophages cells. Overall, our data indicate<br />

important functional roles <strong>of</strong> L. monocy<strong>to</strong>genes Csp proteins in promotion <strong>of</strong> cellular<br />

functions which facilitate environmental stress resistance and host cell invasion<br />

processes <strong>of</strong> this food-borne pathogen.<br />

Different levels <strong>of</strong> flagellin detected during growth<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes strains at low temperature<br />

Cabrita, P. 1,2,3*, Batista1, S., Moes, S. 4, Jenö, P. 4, Trigo, M. J. 3, Boavida Ferreira, R. 1,2<br />

and Bri<strong>to</strong>, L. 1<br />

1. CBAA/ Departamen<strong>to</strong> de Botânica e Engenharia Biológica, Institu<strong>to</strong> Superior de Agronomia,<br />

Technical University <strong>of</strong> Lisbon, Lisbon, Portugal<br />

2. Institu<strong>to</strong> de Tecnologia Química e Biológica, New University <strong>of</strong> Lisbon, Oeiras, Portugal<br />

3. Institu<strong>to</strong> Nacional dos Recursos Biológicos, IP, Oeiras, Portugal<br />

4. Department <strong>of</strong> Biochemistry, Biozentrum <strong>of</strong> the University <strong>of</strong> Basel, Basel, Switzerland<br />

Listeria monocy<strong>to</strong>genes can <strong>to</strong>lerate a wide range <strong>of</strong> environmental stresses.<br />

Growth at low temperatures is a stress that L. monocy<strong>to</strong>genes <strong>of</strong>ten has <strong>to</strong> face<br />

since refrigeration is present throughout the food chain. This study aimed <strong>to</strong> investigate<br />

whether unique or common proteins are up- or down-regulated under<br />

nutritional stress conditions and low temperature. To achieve this, a simplified<br />

methodology <strong>to</strong> analyse the extracellular protein pr<strong>of</strong>iles <strong>of</strong> four L. monocy<strong>to</strong>genes<br />

strains was used. These strains, belonging <strong>to</strong> four different serovars, were selected<br />

according <strong>to</strong> differences in virulence. Cultures were incubated in minimal medium<br />

(Modified Welshimer Broth) at 10 °C. Proteins present in the supernatants <strong>of</strong> the<br />

cell cultures, in late exponential phase, were precipitated and separated by SDS-<br />

PAGE, using equivalent amounts <strong>of</strong> <strong>to</strong>tal secreted proteins. For each bacterial<br />

strain, at least three independent culture assays were set. The most abundant<br />

polypeptide bands and those suggesting the widest range in differential expression<br />

among strains were identified by ESI LC / MS-MS. The p60 virulence protein,<br />

one <strong>of</strong> the major proteins previously detected at 37 °C, was still detected in<br />

relatively high amounts at 10 °C in all strains. The virulence proteins listeriolysin<br />

O and internalin C expressed at 37 °C, were not detected at 10 °C in all strains, confirming<br />

previous findings. In serovar 1/2a strain, flagellin was the major protein<br />

identified. Two other strains, from the rare serovares 4c and 4d/4e, showed lower<br />

levels <strong>of</strong> flagellin w<strong>here</strong>as for the serovar 4b strain no flagellin was detected. Flagella<br />

have been shown <strong>to</strong> act as media<strong>to</strong>rs in bacteria attachment <strong>to</strong> stainless steel<br />

surfaces. Serogroup 1/2a has been reported as prevalent among food isolates.<br />

T<strong>here</strong>fore, the different levels <strong>of</strong> flagellin detected may be related with the ability<br />

<strong>of</strong> some strains <strong>to</strong> colonize food contact surfaces at refrigeration temperatures.<br />

This hypothesis will be further investigated.<br />

* Participation Supported by IUFoST<br />

R E F E R E N C E<br />

A / O<br />

01<br />

R E F E R E N C E<br />

A / O<br />

02<br />

33


R E F E R E N C E<br />

A / O<br />

03<br />

R E F E R E N C E<br />

A / O<br />

04<br />

34<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

Phenotypic and corresponding transcrip<strong>to</strong>mic responses <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes strains in the presence <strong>of</strong> unpro<strong>to</strong>nated organic acids<br />

Lee Chang, K. J., Pinfold, T., Koshy, A. and Bowman, J. P.<br />

University <strong>of</strong> Tasmania, Australia<br />

Sodium diacetate-resistant L. monocy<strong>to</strong>genes food and clinical isolates delineated<br />

through a culture-based screening process were found <strong>to</strong> possess significantly better<br />

survival levels following challenge <strong>to</strong> pH 2.4 acid challenges compared <strong>to</strong> various<br />

reference strains. Sodium diacetate was found <strong>to</strong> directly and substantially<br />

improve survival and was synergistic with acid <strong>to</strong>lerance resultant from the shift <strong>to</strong><br />

the stationary growth phase. The resistant strains had comparatively lower intracellular<br />

levels <strong>of</strong> acetate and K + ions that corresponded <strong>to</strong> lower transmembrane<br />

delta pH values. Another observed feature was that when grown in the presence <strong>of</strong><br />

sodium diacatete cell wall stability was enhanced as revealed by lysis assays, utilising<br />

bead-beating and mutanolysin, suggesting acetate induces various cell wall<br />

modifications that may also influence diffusion <strong>of</strong> unpro<strong>to</strong>nated organic acid in<strong>to</strong><br />

cells. Transcrip<strong>to</strong>mic analysis <strong>of</strong> pH 5.0-habituated sodium diacetate-resistant<br />

strain FW04/0025 compared with reference strain EGD that were grown with<br />

21 mM (8 mM unpro<strong>to</strong>nated) sodium diacetate at pH 5.0 indicated substantial variation<br />

in genetic responses with EGD much more reactionary <strong>to</strong> the presence <strong>of</strong><br />

mineral acid. These differences were reflected in regulon-level gene expression<br />

trends with EGD strongly activating and repressing the SigB- and CodY regulons,<br />

respectively, while in FW04/0025 the response <strong>of</strong> these regulons were muted. The<br />

transcrip<strong>to</strong>me <strong>of</strong> FW04/0025 only becomes more congruent with that <strong>of</strong> EGD<br />

when in the presence <strong>of</strong> sodium diacetate, though several distinct genetic expression<br />

differences still occur. Gene expression trends deriving from exposure <strong>to</strong><br />

sodium diacetate suggest extensive cell wall and membrane modification, alterations<br />

<strong>to</strong> branched-chain amino acid/fatty acid metabolism/biosynthesis, induction<br />

<strong>of</strong> an SOS-like DNA repair and thioredoxin-mediated antioxidant responses<br />

occurs, though strain-level specific responses are quite divergent. Corresponding<br />

proteomic analyses are underway <strong>to</strong> further characterize responses <strong>to</strong> food preservative<br />

organic acids by L. monocy<strong>to</strong>genes.<br />

Internalin pr<strong>of</strong>iling, multilocus sequence typing<br />

and virulence assesments suggest evolutionary his<strong>to</strong>ry<br />

<strong>of</strong> the Listeria monocy<strong>to</strong>genes-Listeria innocua clade<br />

Chen, J. and Fang, W.<br />

Zhejiang University, China<br />

The morphological, ecological, biochemical and genetic resemblance, and the clear<br />

difference <strong>of</strong> virulence between L. monocy<strong>to</strong>genes and L. innocua make this bacterial<br />

clade attractive as models <strong>to</strong> examine the evolution <strong>of</strong> pathogenicity. This<br />

study was attempted <strong>to</strong> examine the population structure <strong>of</strong> L. monocy<strong>to</strong>genes and<br />

L. innocua, and further <strong>to</strong> investigate the microevolution in this clade via pr<strong>of</strong>iling<br />

<strong>of</strong> 37 internalin genes, MLST analysis <strong>of</strong> gyrB-sigB-dapE-hisJ-ribC-purM-gap-tufbetL<br />

gene cluster, and detection <strong>of</strong> 17 virulence genes, <strong>to</strong>gether with in vitro and in<br />

vivo virulence assessments. Results indicate that L. monocy<strong>to</strong>genes comprises three<br />

recognized lineages I, II and III, including a set <strong>of</strong> lineage III strains displaying<br />

strong phospholipase activity and subdued virulence. While resembling L. monocy<strong>to</strong>genes<br />

in having a nearly identical Listeria pathogenicity island I, and inlA and<br />

inlB, these nonpathogenic L. monocy<strong>to</strong>genes strains harbor notably altered prfA,<br />

actA and plcB, and share many similar internalin gene deletions with L. innocua,<br />

e.g., inlJ, inlC, inlI and inlGHE. On the other hand, L. innocua represents a young<br />

species descending from L. monocy<strong>to</strong>genes, which comprises four subgroups: two<br />

major subgroups I and II, and one atypical subgroup IV exhibiting the least genetic<br />

distance <strong>to</strong> L. monocy<strong>to</strong>genes. All L. innocua strains lack 17 virulence genes<br />

found in L. monocy<strong>to</strong>genes, except for subgroup IV strains harboring inlJ, and are<br />

nonpathogenic <strong>to</strong> mice. As shown by the estimation <strong>of</strong> the time <strong>to</strong> the most recent<br />

common ances<strong>to</strong>r, L. monocy<strong>to</strong>gene lineages I and II appeared at approximately<br />

the same time, and this is also the case with L. innocua subgroups I and II. The<br />

nonpathogenic L. monocy<strong>to</strong>genes strains and L. innocua subgroups IV constitute<br />

the possible evolutionary intermediates between L. monocy<strong>to</strong>genes and L. innocua.<br />

The evolutionary his<strong>to</strong>ry in the L. monocy<strong>to</strong>genes-L. innocua clade represents a<br />

rare example <strong>of</strong> evolution <strong>to</strong>wards reduced virulence <strong>of</strong> pathogens.


The SOS response <strong>of</strong> Listeria monocy<strong>to</strong>genes is involved<br />

in stress resistance, mutagenesis, and bi<strong>of</strong>ilm formation<br />

van der Veen, S. 1,2 and Abee, T. 1,2<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

1. Top Institute Food and Nutrition (TIFN), Wageningen, The Netherlands<br />

2. Labora<strong>to</strong>ry <strong>of</strong> Food Microbiology, Wageningen University and Research Centre, Wageningen,<br />

The Netherlands<br />

The food-borne pathogen Listeria monocy<strong>to</strong>genes is widely distributed in the environment.<br />

As a consequence, raw materials used by the food industry could introduce<br />

L. monocy<strong>to</strong>genes <strong>to</strong> food processing facilities. L. monocy<strong>to</strong>genes has<br />

evolved various strategies and networks <strong>to</strong> survive and adapt <strong>to</strong> changing conditions<br />

e.g. during food processing. One <strong>of</strong> the stress response mechanisms we recently<br />

identified in L. monocy<strong>to</strong>genes is the SOS response.<br />

The SOS response is a conserved inducible pathway that is involved in DNA repair<br />

and restart <strong>of</strong> stalled replication forks. We identified the SOS response regulon <strong>of</strong> L.<br />

monocy<strong>to</strong>genes and showed that it is important for stress resistance and adaptive<br />

mutagenesis. In the present study, we investigated the role <strong>of</strong> the SOS response in<br />

L. monocy<strong>to</strong>genes bi<strong>of</strong>ilm formation. L. monocy<strong>to</strong>genes static bi<strong>of</strong>ilms on polystyrene<br />

and glass consists <strong>of</strong> a homogeneous layer, while on stainless steel L. monocy<strong>to</strong>genes<br />

bi<strong>of</strong>ilms consist <strong>of</strong> single attached cells or microcolonies. Static bi<strong>of</strong>ilms<br />

contain the small rod-shaped morphology, which is very similar <strong>to</strong> the morphology<br />

<strong>of</strong> plank<strong>to</strong>nic cells. However, L. monocy<strong>to</strong>genes continuous flow bi<strong>of</strong>ilms consist <strong>of</strong><br />

ball-shaped microcolonies, which are surrounded by a dense network <strong>of</strong> knitted<br />

chains composed <strong>of</strong> elongated cells. We showed that continuous flow bi<strong>of</strong>ilm formation<br />

and not static bi<strong>of</strong>ilm formation is dependent on the SOS response. Using<br />

Q-PCR analysis, promoter reporters, and SOS response mutants, we showed that<br />

the SOS response is activated during knitted-chain bi<strong>of</strong>ilm formation and that deletion<br />

<strong>of</strong> its regulon member yneA, which is involved in cell elongation during SOS response<br />

activation, results in diminished bi<strong>of</strong>ilm formation in continuous flow conditions.<br />

Furthermore, we demonstrated that activation <strong>of</strong> the SOS response during<br />

continuous flow bi<strong>of</strong>ilm formation induced mutagenesis: wild-type bi<strong>of</strong>ilms<br />

showed considerably higher rifampicin resistant fractions than ∆recA bi<strong>of</strong>ilms or<br />

wild-type plank<strong>to</strong>nic cultures. Our results show that the SOS response <strong>of</strong> L. monocy<strong>to</strong>genes<br />

is important for stress resistance, adaptive mutagenesis, and continuous<br />

flow bi<strong>of</strong>ilm formation, and may t<strong>here</strong>fore contribute <strong>to</strong> the survival and persistence<br />

<strong>of</strong> this pathogen in food processing environments.<br />

Clonal diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes, a worldwide perspective<br />

Chenal-Francisque, V. 1, Lopez, J. 1, Cantinelli, T. 1, Caro, V. 2, Tran, C. 2, Leclerq, A. 1,<br />

Lecuit, M. 1 and Brisse, S. 2<br />

1. Institut Pasteur, National Reference Center and WHO Collaborating Centre for LISTERIA, Paris, France<br />

2. Institut Pasteur, Genotyping <strong>of</strong> Pathogens and Public Health Platform, Paris, France<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen that can cause listeriosis, a severe<br />

invasive disease in human with a high fatality rate. L. monocy<strong>to</strong>genes is widespread<br />

in nature. Molecular typing methods have grouped L. monocy<strong>to</strong>genes in<strong>to</strong> two<br />

major genetic lineages (lineages I and II) and an additional minor lineage (lineage<br />

III). They differ according <strong>to</strong> virulence and ecological origin. We have developed a<br />

DNA sequencing-based subtyping method, multilocus sequence typing (MLST) <strong>to</strong><br />

examine the epidemiology and population genetics <strong>of</strong> L. monocy<strong>to</strong>genes. In the<br />

present study, we have undertaken the first spatiotemporal analysis <strong>of</strong> L. monocy<strong>to</strong>genes<br />

biodiversity by sequencing internal portions <strong>of</strong> seven housekeeping genes<br />

in 300 strains isolated from the six continents. We selected a set <strong>of</strong> 300 unrelated<br />

strains <strong>of</strong> L. monocy<strong>to</strong>genes collected from 41 countries and 6 continents isolated<br />

between 1935 and 2009 from different sources. MLST data based on seven housekeeping<br />

genes (3,288 nucleotides) were obtained as described previously (Ragon et<br />

al., PLoS Pathogens (2008)). The results show a pattern <strong>of</strong> biodiversity similar <strong>to</strong><br />

the one we had initially reported in France. Indeed, the population structure is<br />

similar <strong>to</strong> that obtained on 360 strains isolated principally from France by Ragon<br />

et al. Sequences and allelic pr<strong>of</strong>iles are available on the Internet-accessible database<br />

(www.pasteur.fr/mlst). L. monocy<strong>to</strong>genes appears <strong>to</strong> exhibit a homogeneous<br />

clonal diversity across continents, indicating a high dispersal rate at a global scale.<br />

R E F E R E N C E<br />

A / O<br />

05<br />

R E F E R E N C E<br />

A / O<br />

06<br />

35


R E F E R E N C E<br />

A / O<br />

07<br />

R E F E R E N C E<br />

A / O<br />

08<br />

36<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

Life without a cell wall: Listeria monocy<strong>to</strong>genes L-form cells feature<br />

a unique mode <strong>of</strong> division<br />

Briers, Y., Dell’Era, S., Schuppler, M. and Loessner, M. J.<br />

Institute <strong>of</strong> Food, Nutrition and Health, ETH Zurich, Switzerland<br />

Cell wall-deficient bacteria referred <strong>to</strong> as L-forms have lost the ability <strong>to</strong> maintain<br />

or build a rigid peptidoglycan envelope. Although L-forms had been studied<br />

for decades and many reports exist on their morphological, serological and biochemical<br />

properties, very little was known about the basic cell biology and molecular<br />

mechanisms underlying the transformation process. Of particular interest is<br />

the question how L-form bacteria are able <strong>to</strong> proliferate in the absence <strong>of</strong> a mature<br />

cell wall. We have investigated the biology <strong>of</strong> stable, non-reverting Listeria monocy<strong>to</strong>genes<br />

L-form cells (native and GFP labelled). Transmission electron microscopy<br />

demonstrated that L-form cells are devoid <strong>of</strong> the typical thick Listeria<br />

type cell wall small, and form small pro<strong>to</strong>plast-like vesicles as well as multi-nucleated<br />

macro-cells, surrounded only by a cy<strong>to</strong>plasmic membrane. They lack peptidoglycan-bound<br />

proteins such as Internalin A, w<strong>here</strong>as membrane-anchored<br />

proteins such as Internalin B are still present. Moni<strong>to</strong>ring L-form growth by timelapse<br />

confocal laser scanning microscopy revealed the development and maturation<br />

<strong>of</strong> vesicles within maternal L-form cells. We propose a novel model for growth<br />

and division <strong>of</strong> L-form bacteria, which may explain their ability <strong>to</strong> multiply in the<br />

absence <strong>of</strong> a rigid cell wall. Furthermore, transcrip<strong>to</strong>me analysis <strong>of</strong> parental and Lform<br />

L. monocy<strong>to</strong>genes was performed using whole genome hybridization arrays.<br />

Compared <strong>to</strong> parental bacteria, L-forms feature downregulated metabolic functions<br />

correlating with the dramatic shift in surface <strong>to</strong> volume ratio, w<strong>here</strong>as upregulation<br />

<strong>of</strong> stress genes reflects the difficulties in adapting <strong>to</strong> this unusual, cellwall<br />

deficient lifestyle. We also observed that L-form cells taken up in<strong>to</strong><br />

macrophages are not killed but seem able <strong>to</strong> survive for prolonged periods <strong>of</strong> at<br />

least 48 hours. In conclusion, we show that L. monocy<strong>to</strong>genes L-forms (i) can arise<br />

and survive in the environment, (ii) are able <strong>to</strong> multiply and divide, and (iii) show<br />

intracellular survival in macrophages.<br />

Pangenomic analysis <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Deng, X. 1, Phillippy, A. M. 2, Li, Z. 1, Salzberg, S. L. 2, Tor<strong>to</strong>rello, M. L. 3 and Zhang, W. 1<br />

1. National Center for Food Safety and Technology, Illinois Institute <strong>of</strong> Technology, Summit, USA<br />

2. Center for Bioinformatics and Computational Biology, University <strong>of</strong> Maryland, College Park, USA<br />

3. National Center for Food Safety and Technology, Food and Drug Administration, Summit, USA<br />

Listeria monocy<strong>to</strong>genes is well known for its adaptability <strong>to</strong> diverse environment<br />

and host niches and its high fatality rate among infected immunocompromised<br />

populations. Three genetic lineages have been identified in this species. Strains<br />

<strong>of</strong> genetic lineages I and II account for >90 % <strong>of</strong> human infections in the United<br />

States, w<strong>here</strong>as strains from genetic lineage III are rarely implicated in human<br />

infections for unclear reasons. Here we compare the genomes <strong>of</strong> 26 L. monocy<strong>to</strong>genes<br />

strains representing the three lineages based on both in silico comparative<br />

genomic analysis and high-density, pan-genomic DNA array hybridizations. We<br />

uncover 86 genes and 8 small regula<strong>to</strong>ry RNAs that likely make L. monocy<strong>to</strong>genes<br />

lineages differ in carbohydrate utilization and stress resistance during their residence<br />

in natural habitats and passage through the host gastrointestinal tract. We<br />

also identify 2,330 <strong>to</strong> 2,456 core genes in the listerial pan-genome that define this<br />

species and assess the impact <strong>of</strong> lysogenic bacteriophages on genomic diversification.<br />

Phylogenomic reconstructions based on 3,560 homologous groups suggest<br />

a polyphyletic population infrastructure and gradual loss <strong>of</strong> genes as this saprophytic<br />

species diversified in<strong>to</strong> a rare and probably defective lineage.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

Evidence for an antiporter-independent glutamate decarboxylase<br />

(GAD) system in Listeria monocy<strong>to</strong>genes: Influence <strong>of</strong> growth media<br />

on GAD system activity<br />

Karatzas, K.-A., Brennan, O., Heavin, S. and O’Byrne, C. P.<br />

SFI, Ireland<br />

The GAD system plays an important role in survival <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

under acidic conditions (eg. acidic foods) and in virulence (eg. survival in s<strong>to</strong>mach).<br />

The GAD system imports extracellular L-glutamate (Glu) e through an antiporter,<br />

converts it <strong>to</strong> γ-aminobutyric acid (GABA) resulting in the consumption<br />

<strong>of</strong> an intracellular pro<strong>to</strong>n and thus the increase <strong>of</strong> the intracellular pH. We show<br />

for the first time that (Glu) e added in Defined Medium (DM) is not used by the<br />

GAD system and t<strong>here</strong>fore it cannot increase the ability <strong>of</strong> this bacterium <strong>to</strong> grow<br />

in mild acidic conditions, neither does it enhance acid resistance at lethal pH values<br />

as it does in BHI. The rate <strong>of</strong> GABA export was moni<strong>to</strong>red in BHI at various pH<br />

values showing that it initiates at pH 4.6 and the rate increases as pH values decrease.<br />

We also demonstrate that t<strong>here</strong> are activa<strong>to</strong>rs <strong>of</strong> the antiporter-based GAD<br />

system in BHI, while its inactivity in DM is not due <strong>to</strong> the presence <strong>of</strong> inhibi<strong>to</strong>rs in<br />

this medium. Activation was at the transcription level with the expression <strong>of</strong><br />

gadD2T2 being more than 100-fold higher in BHI than in DM w<strong>here</strong> levels <strong>of</strong><br />

gadD2 were undetectable. Furthermore we demonstrated for first time that in<br />

both acidified DM and BHI, L. monocy<strong>to</strong>genes accumulates high levels <strong>of</strong> intracellular<br />

GABA (GABA) i (> 42 mM) and despite the slower rate <strong>of</strong> accumulation in DM<br />

the final levels were identical in both media. Since DM does not contain any (Glu) e<br />

we have shown for first time that this bacterium converts (Glu) i <strong>to</strong> (GABA) i , which<br />

is not exported and thus is s<strong>to</strong>red intracellularly. This suggests an alternative<br />

mechanism <strong>of</strong> acid resistance based on the GAD system that circumvents the antiporter-based<br />

mechanism. We suggest that the (GABA) i accumulation might<br />

buffer the intracellular pH and t<strong>here</strong>by contribute in survival under extreme acidic<br />

conditions.<br />

Role <strong>of</strong> Listeria monocy<strong>to</strong>genes tyrosine phosphatases in conferring<br />

listeriophage resistance<br />

Paz, R.-N. 1, Eugster, M. R. 2, Zeiman, E. 1, Loessner, M. J. 2 and Calendar, R. 3<br />

1. Hadassah Hebrew University Medical Center, Israel<br />

2. Institute <strong>of</strong> Food, Nutrition and Health, ETH Zurich, Switzerland<br />

3. Department <strong>of</strong> Molecular and Cell Biology, University <strong>of</strong> California, Berkeley, USA<br />

Protein tyrosine phosphatase (PTP)-like proteins exist in many bacteria and are<br />

segregated in<strong>to</strong> 2 major groups: Low molecular weight, and conventional. These<br />

PTP are suggested <strong>to</strong> be involved in many aspects <strong>of</strong> bacterial physiology including<br />

stress response, DNA binding proteins, virulence and capsule/cell wall production.<br />

By annotation Listeria monocy<strong>to</strong>genes (LM) possesses 2 potential low molecular<br />

weight and 2 conventional PTPs. Although no tyrosine kinases were identified<br />

yet in LM, using Immunoprecipitation (IP) on <strong>to</strong>tal cell lysate and MS plus MS/MS<br />

on the IP products we have identified at least 10 tyrosine phosphorylated proteins.<br />

These proteins vary in their physiological function and were associated with carbohydrate<br />

metabolism, DNA and RNA binding, processing and transcription and<br />

transport. Using LM WT strain 10403S, we have created an in-frame deletion mutant<br />

lacking all 4 PTPs, as well as 4 additional complemented strains harboring<br />

each <strong>of</strong> the PTPs. No major physiological differences were observed between the<br />

WT and the mutant lacking all 4 PTPs. However, the deletion mutant strain was<br />

found <strong>to</strong> be resistant <strong>to</strong> listeriophages A511 and P35, and sensitive <strong>to</strong> other listeriophages<br />

such as U153 and A118. This phage resistance was attributed <strong>to</strong> reduced<br />

attachment <strong>to</strong> the cell wall. Additionally, the mutant lacking all PTPs was found <strong>to</strong><br />

lack N-acetylglucosamine in its teichoic acid. Phage sensitivity was rescued in a<br />

complemented strain harboring a low molecular weight PTP (LMO2540). We also<br />

found that attachment <strong>of</strong> the phages is partially res<strong>to</strong>red by the same PTP and by<br />

one with conventional weight PTP (LMO1800). Thus, it seems that PTPs probably<br />

affect many processes in LM, but mostly resistance <strong>to</strong> listeriophages.<br />

R E F E R E N C E<br />

A / O<br />

09<br />

R E F E R E N C E<br />

A / O<br />

10<br />

37


R E F E R E N C E<br />

A / O<br />

11<br />

R E F E R E N C E<br />

A / O<br />

12<br />

38<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

Thiolomics – the thiol: disulfide redox metabolism<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Ondrusch, N. 1, Gopal, S. 2, Fuss, A. 1, Hagen, N. 1, S<strong>to</strong>ll, R. 1, Aharonowitz, Y. 3 and Kreft, J. 1<br />

1. University <strong>of</strong> Würzburg, Biocenter, Germany<br />

2. Dept. <strong>of</strong> Microbiology , University <strong>of</strong> Mysore, India<br />

3. Dept. <strong>of</strong> Molecular Microbiology and Biotechnology, Tel Aviv University, Israel<br />

The thiol: disulfide redox metabolism (TDRM), found in all living cells, constitutes<br />

a multi-component network. It counteracts oxidative stress, serves <strong>to</strong> maintain<br />

the proper intracellular redox potential, assists in protein folding, is essential<br />

for the formation <strong>of</strong> deoxyribonucleotides for DNA synthesis and helps <strong>to</strong> repair<br />

damaged proteins. The best-characterized biological thiols are the tripeptide glutathione<br />

(GSH) and the small proteins <strong>of</strong> the thioredoxin (Trx) and glutaredoxin<br />

(Grx) family. Oxidized GSH and Trx are recycled by cognate reductases (GSH reductase<br />

– Gor/GshR; thioredoxin reductase – TrxB). A considerable number <strong>of</strong><br />

genes/gene products putatively involved in the TDRM <strong>of</strong> L. monocy<strong>to</strong>genes EGDe<br />

has been identified in the genome sequence. It comprises the unique fused gene<br />

gshF (GSH synthetase), grx, two putative gor/gshR, six members <strong>of</strong> the thioredoxin<br />

family, trxB, class I and class III ribonucleotide reductases (nrdAB and<br />

nrdDG), gpo/gpx, prx, tpx, ohrA/R (de<strong>to</strong>xification <strong>of</strong> organic hydroperoxides), msrA<br />

(methionine sulfoxide reductase ) and several regula<strong>to</strong>rs, e.g. perR (peroxide regulon),<br />

spx & rex (redox-sensing regula<strong>to</strong>rs), fur, zur & mntR (metal uptake), nrdR<br />

and the redox-sensitive chaperone hsp33. We investigated the regulation and function<br />

<strong>of</strong> these genes/gene products in the TDRM in particular and in the physiology<br />

<strong>of</strong> L. monocy<strong>to</strong>genes in general, also with respect <strong>to</strong> infectivity and virulence, by the<br />

following approach: i) construction <strong>of</strong> selected mutants, ii) study <strong>of</strong> their multiplication<br />

in vitro and in vivo, and iii) genome-wide transcription pr<strong>of</strong>iling <strong>of</strong> wild<br />

type and mutants under several in vitro conditions (oxidative and disulfide stress)<br />

and in eukaryotic host cell models. A major result was that mutants defective in the<br />

glutathione/glutaredoxin system showed extensive changes in their transcription<br />

pr<strong>of</strong>iles. Affected were virulence genes, genes for regula<strong>to</strong>rs, transporters, stress<br />

response fac<strong>to</strong>rs and also for metabolic enzymes. These results emphasize the central<br />

role <strong>of</strong> the TDRM, their impact on cellular processes and pathogen-host interaction<br />

will be discussed.<br />

RNA-structures acting at a distance<br />

Johansson, J.<br />

Umeå University, Sweden<br />

Riboswitches are RNA-elements acting in cis, controlling expression <strong>of</strong> their<br />

downstream genes through a metabolite-induced alteration <strong>of</strong> their secondary<br />

structure. Here, we demonstrate that an S-adenosylmethionine (SAM) riboswitch,<br />

SreA, in Listeria monocy<strong>to</strong>genes can also function in trans, and act as a non-coding<br />

RNA. We show that SreA controls expression <strong>of</strong> the virulence regula<strong>to</strong>r PrfA by<br />

binding <strong>to</strong> the 5´-untranslated region <strong>of</strong> its mRNA in a temperature-dependent<br />

manner. Absence <strong>of</strong> the SAM riboswitch increases the level <strong>of</strong> PrfA. Thus, the impact<br />

<strong>of</strong> the SAM riboswitch on PrfA highlights a link between virulence and the<br />

nutritional status <strong>of</strong> the bacterium. Together, our results describe a novel role for<br />

riboswitches and a new class <strong>of</strong> regula<strong>to</strong>ry non-coding RNAs in bacteria.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 01–13<br />

Deep RNA sequencing <strong>of</strong> Listeria monocy<strong>to</strong>genes reveals overlapping<br />

and extensive stationary phase and sigma B-dependent transcrip<strong>to</strong>mes,<br />

including multiple highly transcribed noncoding RNAs<br />

Oliver, H. F. 1, Orsi, R. H. 1, Ponnala, L. 1, Keich, U. 2, Wang, W. 1, Sun, Q. 1, Cartinhour, S. 3,<br />

Filiatrault, M. J. 3, Wiedmann, M. 1 and Boor, K. J. 1<br />

1. Cornell University, USA<br />

2. University <strong>of</strong> Sydney, Australia<br />

3. United States Department <strong>of</strong> Agriculture-Agricultural Research Service, Robert W. Holley Center for<br />

Agriculture and Health, USA<br />

Identification <strong>of</strong> specific genes and gene expression patterns important for Listeria<br />

monocy<strong>to</strong>genes survival, transmission and pathogenesis is critically needed <strong>to</strong><br />

enable development <strong>of</strong> more effective control strategies. The stationary phase<br />

stress response transcrip<strong>to</strong>me, which includes many sigma B-dependent genes,<br />

was defined in L. monocy<strong>to</strong>genes using RNA sequencing (RNA-Seq) with the Illumina<br />

Genome Analyzer. Specifically, transcrip<strong>to</strong>mes were compared between stationary<br />

phase cells <strong>of</strong> L. monocy<strong>to</strong>genes 10403S and an otherwise isogenic ∆sigB<br />

mutant, which does not express the alternative sigma fac<strong>to</strong>r σ B, a major regula<strong>to</strong>r<br />

<strong>of</strong> genes contributing <strong>to</strong> stress response, including stresses encountered upon<br />

entry in<strong>to</strong> stationary phase. Overall, 83 % <strong>of</strong> all L. monocy<strong>to</strong>genes genes were transcribed<br />

in stationary phase cells; 42 % <strong>of</strong> currently annotated L. monocy<strong>to</strong>genes<br />

genes showed medium <strong>to</strong> high transcript levels under these conditions. A <strong>to</strong>tal <strong>of</strong><br />

96 genes had significantly higher transcript levels in 10403S than in ∆sigB, indicating<br />

σ B-dependent transcription <strong>of</strong> these genes. RNA-Seq analyses indicate that<br />

a <strong>to</strong>tal <strong>of</strong> 67 noncoding RNA molecules (ncRNAs) are transcribed in stationary<br />

phase L. monocy<strong>to</strong>genes, including 7 previously unrecognized putative ncRNAs.<br />

Application <strong>of</strong> a dynamically trained Hidden Markov Model, in combination with<br />

RNA-Seq data, identified 65 putative σ B promoters upstream <strong>of</strong> 82 <strong>of</strong> the 96 σ Bdependent<br />

genes and upstream <strong>of</strong> the one σ B-dependent ncRNA. The RNA-Seq<br />

data also enabled annotation <strong>of</strong> putative operons as well as visualization <strong>of</strong> 5’- and<br />

3’-UTR regions. The results from these studies provide powerful evidence that<br />

RNA-Seq data combined with appropriate bioinformatics <strong>to</strong>ols allow quantitative<br />

characterization <strong>of</strong> L. monocy<strong>to</strong>genes transcrip<strong>to</strong>mes, thus providing exciting new<br />

strategies for exploring transcriptional regula<strong>to</strong>ry networks.<br />

R E F E R E N C E<br />

A / O<br />

13<br />

39


R E F E R E N C E<br />

B / O<br />

14<br />

R E F E R E N C E<br />

B / O<br />

15<br />

40<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

ORAL PRESENTATIONS // O / 14–23<br />

Probiotics reduce Listeria monocy<strong>to</strong>genes-induced tissue invasion and<br />

stillbirths in pregnant guinea pigs<br />

Smith, M. A., Agyekum, K. and Williams, D.<br />

Environmental Health Science Department, University <strong>of</strong> Geórgia, USA<br />

One-third <strong>of</strong> listeriosis cases are pregnancy-related and can lead <strong>to</strong> miscarriage<br />

or stillbirth, premature delivery, or infection <strong>of</strong> the newborn. Recently we have<br />

published a risk assessment based on dose response data from nonhuman primates<br />

and guinea pigs. Both animal models estimate LD 50 s <strong>of</strong> approximately 10 7 L.<br />

monocy<strong>to</strong>genes cfu similar <strong>to</strong> the FAO/ WHO estimated human LD 50 <strong>of</strong> 1.9 x 10 6<br />

cfu based on outbreak data. The similarities between the dose response curves<br />

and LD 50 s suggest these animal models are appropriate for testing therapies and<br />

preventive strategies applicable <strong>to</strong> humans. Recently, probiotics have been proposed<br />

<strong>to</strong> protect hosts from pathogens introduced <strong>to</strong> the system through ingestion.<br />

Our objective was <strong>to</strong> determine the efficacy <strong>of</strong> probiotics in yogurt in preventing<br />

L. monocy<strong>to</strong>genes invasion and stillbirths. Using our pregnant guinea pig<br />

model for listeriosis, 5 ml <strong>of</strong> yogurt containing Lac<strong>to</strong>bacillus and Bifidobacterium<br />

was administered orally <strong>to</strong> pregnant guinea pigs on gestation days (gd) 32–36. On<br />

gd 35, guinea pigs were orally fed L. monocy<strong>to</strong>genes (10 9 cfu) in sterilized whipping<br />

cream at four hrs after yogurt feeding. By gd 56 in those animals receiving only L.<br />

monocy<strong>to</strong>genes, the pathogen was isolated from 100 %, 75 %, 64 %, 71 % and 71 %<br />

<strong>of</strong> maternal livers and spleens, placentas, fetal livers and brains, respectively. However<br />

in those that received yogurt and L. monocy<strong>to</strong>genes, the pathogen was isolated<br />

from 56 %, 14 %, 17 %, 17 % and 22 % <strong>of</strong> maternal livers and spleens, placentas,<br />

fetal livers and brains respectively. Also, 75 % <strong>of</strong> pregnant guinea pigs treated with<br />

10 8 L. monocy<strong>to</strong>genes cfu have stillbirths compared <strong>to</strong> 14 % when treated with yogurt<br />

and L. monocy<strong>to</strong>genes. These results present opportunities <strong>to</strong> develop treatment<br />

and preventive therapies for listeriosis. In summary, consumption <strong>of</strong> yogurt<br />

containing Bifidobacterium and Lac<strong>to</strong>bacillus reduced the invasion and number<br />

<strong>of</strong> stillbirths after L. monocy<strong>to</strong>genes exposure in pregnant guinea pigs.<br />

Listeriolysin O favors Listeria monocy<strong>to</strong>genes growth in co-culture with<br />

the ciliate Tetrahymena pyriformis<br />

Ermolaeva, S. and Pushkareva, V.<br />

Gamaleya Institute <strong>of</strong> Epidemiology and Microbiology, Russian Federation<br />

Listeria monocy<strong>to</strong>genes was isolated from soil, water, sewage and sludge. We explored<br />

the potential <strong>of</strong> L. monocy<strong>to</strong>genes major virulence fac<strong>to</strong>r Listeriolysin O<br />

(LLO) <strong>to</strong> promote interactions between L. monocy<strong>to</strong>genes and the ubiqui<strong>to</strong>us inhabitant<br />

<strong>of</strong> natural ecosystems bacteriovorous free-living ciliate Tetrahymena pyriformis.<br />

Axenic T. pyriformis and the following bacterial strains were used: wild<br />

type L. monocy<strong>to</strong>genes strains EGDe, VIMVR081, VIMVW039, VIMHA034,<br />

VIMVF870, EGDe derivative EGDe::Dhly and wild type L. innocua strain<br />

NCTC11288. Experiments were performed in LB broth at 28 °C. Bacteria were<br />

counted bacteriologically or by qPCR. T. pyriformis trophozoites and cysts were<br />

counted using light microscopy. The hly gene was introduced in<strong>to</strong> pTRKH2 vec<strong>to</strong>r,<br />

and the same plasmid supplemented with prfA* gene was used <strong>to</strong> express LLO in L.<br />

innocua. Guinea pigs were infected intraconjunctivally or per os with L. monocy<strong>to</strong>genes<br />

culture or with T. pyriformis cysts infected L. monocy<strong>to</strong>genes. After 7 days<br />

<strong>of</strong> co-culturing, wild type L. monocy<strong>to</strong>genes strains reduced trophozoite and increased<br />

cyst concentrations up <strong>to</strong> 21.1 and 6.1 times, respectively. EGDe::Dhly<br />

failed <strong>to</strong> cause mortality among pro<strong>to</strong>zoa and <strong>to</strong> trigger pro<strong>to</strong>zoan encystment. In<br />

concordance, LLO deficiency deteriorated L. monocy<strong>to</strong>genes growth in the presence<br />

<strong>of</strong> T. pyriformis. Replenishment <strong>of</strong> the hly gene in the mutant strain res<strong>to</strong>red<br />

<strong>to</strong>xicity <strong>to</strong>wards pro<strong>to</strong>zoa. L. innocua transformed with the LLO-expressing plasmid<br />

caused extensive mortality and encystment in ciliates. L. monocy<strong>to</strong>genes EGDe<br />

entrapped in cysts caused infection in guinea pigs upon ocular and oral infection.<br />

The L. monocy<strong>to</strong>genes virulence fac<strong>to</strong>r LLO promotes bacterial survival and is responsible<br />

for L. monocy<strong>to</strong>genes <strong>to</strong>xicity for pro<strong>to</strong>zoa and induction <strong>of</strong> pro<strong>to</strong>zoan<br />

encystment. T<strong>here</strong>fore, LLO activity might support bacterial survival in the natural<br />

habitat outside <strong>of</strong> a host.


A<strong>to</strong>py is a risk fac<strong>to</strong>r for listeriosis<br />

Kawamo<strong>to</strong>, K., Matsubara, S., Da Silva, M. and Makino, S.-I.<br />

Obihiro Univ. Agri. Vet. Med., Japan<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

ORAL PRESENTATIONS // O / 14–23<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive bacterium that causes meningitis, bacteremia,<br />

and febrile gastroenteritis. The outcome <strong>of</strong> listeriosis is dependent on<br />

host fac<strong>to</strong>rs such as age, pregnancy, and HIV infection, which influence host immunocompetency.<br />

A Th-1 type cy<strong>to</strong>kine IFN-gamma; plays an important role in<br />

the innate immune response against intracellular bacterial pathogens. In contrast,<br />

Th2-biased immune responses <strong>of</strong>ten associate with most a<strong>to</strong>pic diseases. In this<br />

study, we examined whether a<strong>to</strong>py affected the immune response <strong>to</strong> L. monocy<strong>to</strong>genes<br />

infection. NC/Nga is a model mouse for human a<strong>to</strong>pic dermatitis, which has<br />

a genetic predisposition <strong>to</strong> develop a<strong>to</strong>pic skin lesion. We compared the susceptibility<br />

<strong>to</strong> listeriosis <strong>of</strong> NC/Nga mice with BALB/c and C57BL/6 mice. The course <strong>of</strong><br />

infection was characterized by moni<strong>to</strong>ring survival <strong>of</strong> mice, and determination <strong>of</strong><br />

bacterial numbers in the organs. NC/Nga mice were highly susceptible <strong>to</strong> L. monocy<strong>to</strong>genes<br />

infection: the 50 % lethal dose was significantly lower and the number <strong>of</strong><br />

bacteria in livers, spleens and brains were higher in NC/Nga than those for other<br />

inbred mice. The increased permeability <strong>of</strong> blood-brain barrier was observed in<br />

NC/Nga brains at day 3 post infection, but not in BALB/c and C57BL/6. As compared<br />

<strong>to</strong> BALB/c and C57BL/6 mice, plasma IFN-gamma-levels <strong>of</strong> NC/Nga were<br />

comparative. However, markedly increased IL-10 levels were detected in NC/Nga<br />

plasma. Pretreatment with neutralizing antibodies <strong>to</strong> IL-10 partially protected<br />

mice but retarded the clearance <strong>of</strong> bacteria. Yet the molecular mechanisms by<br />

which the infection <strong>of</strong> L. monocy<strong>to</strong>genes induces overproduction <strong>of</strong> IL-10 in<br />

NC/Nga mice remain unknown, our results suggest that the differential cy<strong>to</strong>kine<br />

production may at least partially underlie the higher susceptibility <strong>to</strong> L. monocy<strong>to</strong>genes<br />

in NC/Nga mice. Considering the increased prevalence <strong>of</strong> a<strong>to</strong>pic diseases,<br />

an a<strong>to</strong>pic phenotype may be a potential risk fac<strong>to</strong>r for listeriosis.<br />

Cancer immunotherapy using novel Listeria monocy<strong>to</strong>genes-bacterial<br />

vec<strong>to</strong>rs <strong>to</strong> target the vasculature <strong>of</strong> progressive tumors<br />

Paterson, Y., Seavey, M. M., Maciag, P. C. and Sewell, D.<br />

University <strong>of</strong> Pennsylvania, USA<br />

For nearly 20 years our labora<strong>to</strong>ry has been developing Listeria monocy<strong>to</strong>genes<br />

(Lm) as a vaccine carrier <strong>to</strong> introduce tumor and viral antigens <strong>to</strong> the immune<br />

system for the immunotherapy <strong>of</strong> cancer and as prophylactic vaccines for infectious<br />

disease. Given the antigenic instability <strong>of</strong> tumor cells, we recently turned<br />

our attention <strong>to</strong> the use <strong>of</strong> immunotherapy <strong>to</strong> destroy cells actively involved in<br />

forming new blood vessels that support the growth and spread <strong>of</strong> cancer. We constructed<br />

Lm expression systems that would target two central cell types involved<br />

in angiogenesis – endothelial cells and pericytes. Two proteins are highly expressed<br />

on each cell during active angiogenesis – Vascular Endothelial Growth<br />

Fac<strong>to</strong>r Recep<strong>to</strong>r-2 (VEGFR2) (endothelial cells) and High Molecular Weight<br />

Melanoma Associated Antigen (HMWMAA) (pericytes). We selected fragments<br />

from each molecule that included peptides predicted <strong>to</strong> bind <strong>to</strong> the HLA-A2 molecule.<br />

We fused the genes encoding these fragments <strong>to</strong> the gene encoding the first<br />

420 residues <strong>of</strong> Listeriolysin-O (LLO) and used Lm <strong>to</strong> deliver these fusion proteins.<br />

Even though the immune system should be <strong>to</strong>lerant <strong>to</strong> these self-molecules,<br />

both vaccines elicited potent anti-tumor CTL responses. Lm-LLO-VEGFR2 was<br />

able <strong>to</strong> reduce tumor microvascular density (MVD), cause regression <strong>of</strong> established<br />

breast tumors and protect against tumor re-challenge 100+ days post last<br />

immunization. The Lm-LLO-HMWMAA vaccine also dramatically reduced MVD,<br />

induced regression <strong>of</strong> established breast tumors, increased CD8+ T cell infiltration,<br />

and, in addition, reduced the pericyte coverage <strong>of</strong> intra-tumoral blood vessels.<br />

Interestingly anti-tumor efficacy was dependent on epi<strong>to</strong>pe spreading <strong>to</strong> the<br />

tumor-associated antigen Her-2/neu. Neither immunotherapeutic interfered with<br />

the generation <strong>of</strong> normal vasculature in wound healing or gestation. We are currently<br />

testing these vaccines <strong>to</strong> impact the progression and metastasis <strong>of</strong> advanced<br />

breast cancer, we hypothesize that reduced angiogenesis will prevent the formation<br />

<strong>of</strong> distal metastases blunting cancer spread and improving overall survival.<br />

R E F E R E N C E<br />

B / O<br />

16<br />

R E F E R E N C E<br />

B / O<br />

17<br />

41


R E F E R E N C E<br />

B / O<br />

18<br />

R E F E R E N C E<br />

B / O<br />

19<br />

42<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

ORAL PRESENTATIONS // O / 14–23<br />

The intracellular carbon metabolism <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in comparison <strong>to</strong> that <strong>of</strong> other intracellular bacterial pathogens<br />

replicating in mammalian host cells<br />

Gotz, A. 1, Eylert, E. 2, S<strong>to</strong>ll, R. 1, Eisenreich, W. 2, and Goebel, W. 3<br />

1. Biocenter-Microbiology, University <strong>of</strong> Würzburg, Germany<br />

2. Institute <strong>of</strong> Biochemistry, LMU München, Germany<br />

3. Max-von-Pettenk<strong>of</strong>er Institute, LMU München, Germany<br />

Intracellular bacterial pathogens replicate either in the cy<strong>to</strong>sol (e.g. Listeria monocy<strong>to</strong>genes,<br />

Shigella spp. and the closely related enteroinvasive Escherichia coli<br />

[EIEC]) or in specialized phagosomal compartments (e.g. Salmonella enterica<br />

Serovar Typhimurium) <strong>of</strong> the infected host cells. Our knowledge on the metabolic<br />

adaptation processes between intracellular pathogens and their host cells, allowing<br />

their efficient intracellular growth, and on the influence <strong>of</strong> the intracellular<br />

bacterial metabolism on the expression <strong>of</strong> those virulence genes, that are decisive<br />

for the intracellular life cycle <strong>of</strong> the intracellular pathogens, is still rather limited.<br />

We have carried out initial studies concerning these important questions with the<br />

above mentioned intracellular bacteria. For this goal we constructed mutants impaired<br />

in the uptake and/or catabolism <strong>of</strong> specific carbon sources and studied by<br />

NMR- or MS-based 13C-iso<strong>to</strong>pologue pr<strong>of</strong>iling their intracellular carbon metabolism<br />

in comparison <strong>to</strong> the isogenic wild-type strains after infection <strong>of</strong> suitable<br />

mammalian host cells. The results obtained show that the intracellular carbon<br />

metabolism <strong>of</strong> L. monocy<strong>to</strong>genes differs significantly from that <strong>of</strong> the two other<br />

pathogens. W<strong>here</strong>as glucose, but not glucose-6-P is the major carbon substrate<br />

for intracellular growth <strong>of</strong> EIEC and S. typhimurium, L. monocy<strong>to</strong>genes uses C 3 -<br />

substrates (e.g. glycerol) as major and glucose-6-P as subsidiary carbon sources.<br />

The reason for this difference is apparently the lack <strong>of</strong> high affinity glucose transporters<br />

in L. monocy<strong>to</strong>genes. However, the intracellular C-metabolism <strong>of</strong> all three<br />

pathogens in mammalian host cells is surprisingly flexible. Mutants defective in<br />

the uptake <strong>of</strong> the preferential carbon source switch readily <strong>to</strong> alternative carbon<br />

sources and the lower energy supply <strong>of</strong> the utilized secondary carbon sources<br />

seems <strong>to</strong> be compensated by an increased uptake <strong>of</strong> anabolic monomers from the<br />

host cells. As shown for L. monocy<strong>to</strong>genes the intracellular carbon metabolism appears<br />

<strong>to</strong> be adapted <strong>to</strong> optimal expression <strong>of</strong> the virulence genes that are essential<br />

for the intracellular life style.<br />

LIMP2 links late phagosomal trafficking with the onset<br />

<strong>of</strong> the innate immune response <strong>to</strong> Listeria monocy<strong>to</strong>genes:<br />

a role in macrophage activation<br />

Fernandez-Prie<strong>to</strong>, L., Carrasco-Marin, E., Madrazo-Toca, F., Rodriguez-Del Rio, E.,<br />

Carranza-Cereceda, C. and Alvarez-Dominguez, C.<br />

Immunology Departament, Hospital Santa Cruz de Liencres and Fundacion Marques de<br />

Valdecilla – IFIMAV, Spain<br />

The innate immune response <strong>to</strong> Listeria monocy<strong>to</strong>genes depends on phagosomal<br />

bacterial degradation by macrophages. Here, we describe the role <strong>of</strong> LIMP2, a lysosomal<br />

type III transmembrane glycoprotein and scavenger-like protein, in Listeria<br />

phagocy<strong>to</strong>sis. We show <strong>here</strong> that LIMP2 is not involved in bacterial recognition at<br />

the cell surface but participates in the degradation <strong>of</strong> Listeria within phagosomes.<br />

LIMP2 appears linked <strong>to</strong> Rab5a activation and controls the late-endosomal/lysosomal<br />

fusion machinery. Importantly, LIMP2 deficient mice display a<br />

macrophage-related defect in innate immunity. They produce less acute-phase<br />

pro-inflamma<strong>to</strong>ry cy<strong>to</strong>kines/chemokines, MCP-1, TNF-α and IL-6, but normal<br />

levels <strong>of</strong> IL-12, IL-10 and IFN-γ and a 20-fold increase in susceptibility <strong>to</strong> Listeria<br />

infection. This macrophage dysfunction results in a low listericidal potential, impaired<br />

phago-lysosome transformation in<strong>to</strong> antigen-processing compartments<br />

and uncontrolled LM cy<strong>to</strong>solic growth that fails <strong>to</strong> induce normal levels <strong>of</strong> acutephase<br />

pro-inflamma<strong>to</strong>ry cy<strong>to</strong>kines. T<strong>here</strong>fore, the role <strong>of</strong> LIMP2 appears <strong>to</strong> be<br />

connected <strong>to</strong> the activation <strong>of</strong> Listeria-primed macrophages through internal signals<br />

linking the regulation <strong>of</strong> late trafficking with the onset <strong>of</strong> the innate Listeria<br />

immune response.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

ORAL PRESENTATIONS // O / 14–23<br />

The tetraspanin CD81 is required for entry <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in mammalian cells<br />

Tham, T. N. 1,2,3, Gouin, E. 1,2,3, Rubinstein, E. 4,5, Boucheix, C. 4,5, Cossart, P. 1,2,3<br />

and Pizarro-Cerdá, J. 1,2,3<br />

1. Institut Pasteur, Unité des Interactions Bactéries-Cellules, Département de Biologie Cellulaire<br />

et Infection, Paris, France<br />

2. INSERM, U604, Paris, France<br />

3. INRA, USC2020, Paris, France<br />

4. Institut André Lw<strong>of</strong>f, Université Paris-Sud, Villejuif, France<br />

5. INSERM, U602, Villejuif, France<br />

Listeria monocy<strong>to</strong>genes is a facultative intracellular bacterial pathogen that invades<br />

epithelial cells by subverting signaling cascades associated with its two main<br />

cellular recep<strong>to</strong>rs, E-cadherin and Met. We recently identified the type II phosphatidylinosi<strong>to</strong>l<br />

4-kinases (PI4KIIs) α and β as required for bacterial entry downstream<br />

<strong>of</strong> Met. In this work we investigated whether tetraspanins CD9, CD63 and<br />

CD81, which figure among the few described molecular partners <strong>of</strong> the PI4KIIα,<br />

function as molecular adap<strong>to</strong>rs recruiting the PI4KIIα <strong>to</strong> the bacterial entry site.<br />

We observed by fluorescence microscopy that CD9, CD63 and CD81 are expressed<br />

and detected at the cellular surface and also within intracellular compartments,<br />

particularly in the case <strong>of</strong> CD63. In resting cells, colocalization between these<br />

tetraspanins and the PI4KIIα is only detectable in restricted areas <strong>of</strong> the perinuclear<br />

region. Upon infection with Listeria, endogenous CD9, CD63 and CD81 were<br />

recruited at the bacterial entry site but did not colocalize strictly with endogenous<br />

PI4KIIα. Live cell imaging confirmed that tetraspanins and the PI4KIIα do<br />

not follow the same recruitment dynamics <strong>to</strong> the Listeria entry site. Depletion <strong>of</strong><br />

CD9, CD63 and CD81 levels by small interfering RNA (siRNA) demonstrated that<br />

CD81 is required for bacterial internalization, identifying for the first time a role<br />

for a member <strong>of</strong> the tetraspanin family in the entry <strong>of</strong> Listeria within target cells.<br />

Moreover, depletion <strong>of</strong> CD81 inhibits recruitment <strong>of</strong> PI4KIIα <strong>to</strong> the bacterial entry<br />

site but not that <strong>of</strong> the Met recep<strong>to</strong>r, suggesting that this tetraspanin could act as<br />

a membrane organizer required for the integrity <strong>of</strong> signaling events occurring at<br />

Listeria entry sites.<br />

Listeria innate immune evasion by peptidoglycan modification<br />

Aubry, C.<br />

Department <strong>of</strong> Cellular Biology and Interaction, Institut Pasteur, France<br />

Peptidoglycan (PG) plays an important role in host-pathogen interaction because<br />

PG is the site <strong>of</strong> anchoring <strong>of</strong> virulence fac<strong>to</strong>rs and an important target for the innate<br />

immune system. As reported previously, inactivation <strong>of</strong> pgdA, encoding a PG<br />

N-deacetylase in Listeria monocy<strong>to</strong>genes, revealed a key role <strong>of</strong> PG modification<br />

in virulence as survival <strong>of</strong> the mutant was severely impaired in mice (Boneca et<br />

al., PNAS 2007). Deletion <strong>of</strong> the pgdA gene highly increased sensitivity <strong>to</strong> the bacteriolytic<br />

activity <strong>of</strong> hen egg lysozyme in vitro. The pgdA mutant was rapidly destroyed<br />

within phagosomes and induced a potent pro-inflamma<strong>to</strong>ry response by<br />

macrophages. Interestingly, inactivation <strong>of</strong> the deacetylase induced a strong secretion<br />

<strong>of</strong> IFN beta by infected macrophages, through a surprisingly and uncharacterized<br />

TLR2-dependent pathway. We have now shown that TLR9 but not TLR4<br />

contribute <strong>to</strong> this IFN beta secretion. The key adap<strong>to</strong>rs and the new signaling<br />

pathway leading <strong>to</strong> IFN beta production are under investigation. We found that<br />

L. monocy<strong>to</strong>genes PG is also modified by a putative O-acetyltransferase (OatA) essential<br />

for virulence. A Listeria oatA mutant shows an increased sensitivity <strong>to</strong> hen<br />

egg lysozyme, bacteriocin and cell wall targeting antibiotics in vitro. Its growth is<br />

impaired in macrophages, possibly contributing <strong>to</strong> immune escape. Modification<br />

<strong>of</strong> PG is a highly efficient mechanism used by pathogenic Listeria <strong>to</strong> evade innate<br />

host defenses.<br />

R E F E R E N C E<br />

B / O<br />

20<br />

R E F E R E N C E<br />

B / O<br />

21<br />

43


R E F E R E N C E<br />

B / O<br />

22<br />

R E F E R E N C E<br />

B / O<br />

23<br />

44<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

ORAL PRESENTATIONS // O / 14–23<br />

Constitutive activation <strong>of</strong> the central virulence transcriptional<br />

regula<strong>to</strong>r PrfA enhances Listeria monocy<strong>to</strong>genes pathogenesis but<br />

reduces bacterial fitness outside <strong>of</strong> the host<br />

Freitag, N. E. 1 and Bruno, J. C. 2<br />

1. Department <strong>of</strong> Microbiology & Immunology, University <strong>of</strong> Illinois at Chicago, USA<br />

2. Department <strong>of</strong> Global Health, University <strong>of</strong> Washing<strong>to</strong>n, USA<br />

Listeria monocy<strong>to</strong>genes survives as a saprophyte in soil and decaying vegetation<br />

while maintaining an ability <strong>to</strong> invade mammalian cells and cause serious disease.<br />

Survival and replication <strong>of</strong> L. monocy<strong>to</strong>genes within mammalian hosts requires<br />

the regulated synthesis and expression <strong>of</strong> multiple gene products that enable host<br />

cell invasion, bacterial replication within the cy<strong>to</strong>sol, and spread <strong>to</strong> adjacent cells.<br />

The transcriptional regula<strong>to</strong>r PrfA controls the expression <strong>of</strong> multiple bacterial<br />

virulence fac<strong>to</strong>rs and is required <strong>to</strong> facilitate the L. monocy<strong>to</strong>genes transition from<br />

saprophyte <strong>to</strong> pathogen. PrfA has been shown <strong>to</strong> exist in both low and high activity<br />

states, with the transition <strong>to</strong> high activity occurring within the cy<strong>to</strong>sol <strong>of</strong> host cells.<br />

A number <strong>of</strong> mutations within prfA have been described (prfA* mutations) that<br />

serve <strong>to</strong> lock the protein in<strong>to</strong> a high activity state. We examined the consequences<br />

<strong>of</strong> constitutive PrfA activation on L. monocy<strong>to</strong>genes physiology and pathogenesis<br />

by assessing the fitness <strong>of</strong> prfA* strains in a variety <strong>of</strong> in vitro and in vivo conditions.<br />

prfA* strains exhibit a competitive advantage over wild strains in animal<br />

models <strong>of</strong> infection, with ten-fold higher numbers <strong>of</strong> prfA* bacteria recovered from<br />

the livers and spleens <strong>of</strong> infected mice. However, despite apparently normal<br />

growth in broth culture, the prfA* mutants exhibited a fitness defect when grown<br />

in the presence <strong>of</strong> wild type bacteria, and this fitness defect was exacerbated when<br />

mixed cultures were placed under various stress conditions. Strains containing<br />

prfA* mutations were also defective in their ability <strong>to</strong> adapt <strong>to</strong> a long term survival<br />

phase following prolonged growth in broth culture. Taken <strong>to</strong>gether, these results<br />

indicate that L. monocy<strong>to</strong>genes must maintain a critical balance <strong>of</strong> PrfA activity<br />

<strong>to</strong> promote bacterial fitness both inside and outside <strong>of</strong> infected host cells.<br />

The lvfH gene <strong>of</strong> Listeria monocy<strong>to</strong>genes encodes a novel virulence fac<strong>to</strong>r<br />

highly activated during infection<br />

Carvalho, F., Camejo, A., Ferreira, P., Sousa, S. and Cabanes, D.<br />

IBMC – Institu<strong>to</strong> de Biologia Molecular e Celular, Group <strong>of</strong> Molecular Microbiology,<br />

Universidade do Por<strong>to</strong>, Portugal<br />

Listeria monocy<strong>to</strong>genes is a highly adaptable Gram-positive bacterium that can<br />

induce potentially lethal listeriosis in immunocompromised human hosts. As a<br />

facultative intracellular pathogen, it is able <strong>to</strong> invade and replicate inside different<br />

eukaryotic cell types and, by cell-<strong>to</strong>-cell spread, disseminate infection. Such<br />

processes are highly dependent upon expression <strong>of</strong> specialized proteins which<br />

act as virulence fac<strong>to</strong>rs on certain steps <strong>of</strong> the bacterial infectious cycle. Our recent<br />

studies on the L. monocy<strong>to</strong>genes EGD-e transcription pr<strong>of</strong>ile in infected<br />

mice revealed genes that were highly activated throughout the infection timeline<br />

and could be involved in virulence. Among these is lvfH, a serotype 1/2a-specific<br />

gene that encodes a protein putatively involved in the L-rhamnose biosynthesis<br />

pathway, a mechanism that provides substrate for the decoration <strong>of</strong> cell wall teichoic<br />

acids. In order <strong>to</strong> confirm and characterize the role <strong>of</strong> this gene in L. monocy<strong>to</strong>genes<br />

virulence, we generated an lvfH deletion mutant and investigated the<br />

influence <strong>of</strong> this mutation in the Listeria infectious process. In vitro assays<br />

showed that the mutant was unaffected in its host cell membrane-adhering<br />

properties, but was significantly impaired in its ability <strong>to</strong> invade different mammalian<br />

cell lines. Organs <strong>of</strong> mice intravenously infected with the lvfH mutant<br />

displayed a significant decrease in bacterial load, as compared <strong>to</strong> animals infected<br />

with wild type bacteria. This phenotype is unrelated with an intrinsic<br />

growth defect <strong>of</strong> the mutant, as it presented growth pr<strong>of</strong>iles similar <strong>to</strong> the wild<br />

type strain in broth and within murine macrophages. Animal organs infected<br />

with an lvfH-complemented strain showed wild type-like infection levels, confirming<br />

the specific requirement <strong>of</strong> LvfH for full virulence in this model.<br />

Functional characterization <strong>of</strong> LvfH and analysis <strong>of</strong> the relation between teichoic<br />

acid rhamnosylation and virulence could reveal a new role for teichoic acid<br />

decoration in L. monocy<strong>to</strong>genes pathogenesis.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Human listeriosis due <strong>to</strong> Listeria ivanovii<br />

Guillet, C. 1, Join-Lambert, O. 1, Le Monnier, A. 1, Leclercq, A. 2, Mamzer-Bruneel,<br />

M. F. 1, Bielecka, M. K. 3, Scortti, M. 3, Disson, O. 2,4, Vazquez-Boland, J. 3, Lortholary, O. 1<br />

and Lecuit, M. 1,2,4<br />

1. Necker-Enfants Malades hospital, Paris Descartes University, Paris, France<br />

2. Institut Pasteur, NRC and WHO-CC for Listeria, Paris, France<br />

3. University <strong>of</strong> Edinburgh, Scotland, UK<br />

4. Institut Pasteur, Microbes and host barriers Group, Inserm avenir, Paris, France<br />

The genus Listeria contains two pathogenic species: Listeria monocy<strong>to</strong>genes (Lm)<br />

that infects human and animals, and Listeria ivanovii subps. ivanovii (Lii), considered<br />

a ruminant-specific pathogen. We describe <strong>here</strong> a case <strong>of</strong> gastroenteritis and<br />

bacteremia due <strong>to</strong> Lii in a kidney transplant recipient. A 55-year-old man was referred<br />

<strong>to</strong> our hospital with a three-week his<strong>to</strong>ry <strong>of</strong> non-bloody diarrhoea, vomiting,<br />

dehydratation, and low-grade fever. He underwent a renal transplantation for<br />

chronic renal failure and was chronically infected with hepatitis C virus. S<strong>to</strong>ol and<br />

blood cultures allowed the isolation <strong>of</strong> Lii. Intravenous amoxicillin (6 g/day) and<br />

gentamicin (6 mg/kg/day) treatment was associated with resolution <strong>of</strong> symp<strong>to</strong>ms.<br />

The 4 Lii isolates from this patient had indistinguishable ApaI and SmaI PFGE<br />

patterns, phenotypic characters and classical resistance for antibiotics. These<br />

human isolates were indistinguishable from pro<strong>to</strong>typic ruminant strains based on<br />

(i) the activation status <strong>of</strong> the central virulence gene regula<strong>to</strong>r PrfA, (ii) the presence<br />

<strong>of</strong> the L. ivanovii-specific pathogenicity island LIPI-2 by PCR mapping, and<br />

(iii) invasion assays using Madin-Darby Bovine Kidney cells and huma, HeLa cells.<br />

So, Lii is pathogenic for humans. As for Lm, Lii route <strong>of</strong> infection is foodborne. A<br />

review <strong>of</strong> the literature identified 8 cases, mostly in immunosuppressed patients.<br />

The rarity <strong>of</strong> human Lii infections probably reflects low exposure given the rare<br />

occurrence <strong>of</strong> this species in nature compared <strong>to</strong> Lm.<br />

Foodborne listeriosis in India: An update<br />

Barbuddhe, S. B. 1*, Malik, S. V. S. 2, Ashok Kumar, J. 1, Kalorey, D. R. 3, Kurkure, N. V. 3,<br />

Rawool, D. B. 4, Swain, B. K. 1, Korikanthimath, V. S. 1 and Chakraborty, T. 4<br />

1. ICAR Research Complex for Goa, India<br />

2. Division <strong>of</strong> Veterinary Public Health, Indian Veterinary Research Institute<br />

3. Nagpur Veterinary College, India<br />

4. Institute <strong>of</strong> Medical Microbiology, Justus Liebig University, Giessen, Germany<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen that can cause serious invasive<br />

illness, mainly in certain well-defined high-risk groups, including elderly and immunocompromised<br />

patients, pregnant women, newborns and infants. In India,<br />

the pathogen has been isolated from humans, animals, a variety <strong>of</strong> foods including<br />

milk and milk products, meat and meat products and vegetables. In India, studies<br />

on molecular epidemiological aspects <strong>of</strong> L. monocy<strong>to</strong>genes are largely lacking.<br />

T<strong>here</strong>fore, we do not know the genetic variability <strong>of</strong> strains isolated and their epidemic<br />

potential. The genetic diversity <strong>of</strong> the Listeria strains from various sources<br />

namely, milk and milk products (112), fish and seafood (47), wild life (7), poultry<br />

meat (19), meat and processed meats (37), animal clinical cases (41) and human<br />

clinical cases (21) isolated/collected from different places in the country have been<br />

characterized biochemically and with in vitro assays before attempting for serotyping<br />

and virulence gene pr<strong>of</strong>iles. Out <strong>of</strong> the strains characterized phenotypically,<br />

195 strains have been serotyped using multiplex PCR and pulsed field gel electrophoresis.<br />

The predominant serotype among Indian Listeria isolates is L. monocy<strong>to</strong>genes<br />

4b. Significant variation among the isolates recovered from different<br />

sources has been observed as evident by PFGE analysis. As many as 34 different<br />

PFGE pr<strong>of</strong>iles have been observed. An electronic database for the characterized<br />

strains has been created. This is an interactive web based database so that the data<br />

can be exchanged between labora<strong>to</strong>ries electronically. The molecular characterization<br />

<strong>of</strong> Listeria isolates from India would help in better understanding <strong>of</strong> the<br />

sources <strong>of</strong> infection and their risk assessment, routes <strong>of</strong> transmission <strong>of</strong> the infective<br />

agent, influencing fac<strong>to</strong>rs, clinical forms and virulence characteristic <strong>of</strong><br />

the pathogenic isolates <strong>of</strong> Listeria in order <strong>to</strong> diagnose the infection rapidly and reliably<br />

but also in instituting the effective control measures.<br />

* Participation Supported by Fundação do Oriente<br />

R E F E R E N C E<br />

C / O<br />

24<br />

R E F E R E N C E<br />

C / O<br />

25<br />

45


R E F E R E N C E<br />

C / O<br />

26<br />

R E F E R E N C E<br />

C / O<br />

27<br />

46<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Human listeriosis and co-morbidities in England, 1999 <strong>to</strong> 2008:<br />

quantifying the risk<br />

Mook, P., Grant, K., O’Brien, S. J. and Gillespie, I.<br />

Health Protection Agency, UK<br />

Listeria monocy<strong>to</strong>genes causes a rare but severe food borne disease (listeriosis),<br />

commonly affecting pregnant women, the elderly and the seriously ill. The epidemiology<br />

<strong>of</strong> listeriosis in England and Wales changed between 2001 and 2008,<br />

with more patients aged ≥ 60 years presenting with bacteraemia. We examined<br />

the risks in this age group, and quantified the role <strong>of</strong> co-morbidities for listeriosis<br />

in all age groups. Case-patients were resident in England between 1999 and 2008<br />

and were reported <strong>to</strong> an enhanced national surveillance scheme by hospital microbiologists<br />

using a clinical case questionnaire. We coded co-morbidities <strong>of</strong> non<br />

pregnancy-related cases <strong>of</strong> L. monocy<strong>to</strong>genes infection according <strong>to</strong> ICD-10. These<br />

data were compared with appropriate denomina<strong>to</strong>r data (Hospital Episode Statistics<br />

finished consultant episodes), <strong>to</strong> calculate incidence rates per million consultations<br />

(with appropriate 95 % confidence intervals). Between 1999 and 2008,<br />

1412 non-pregnancy related cases <strong>of</strong> listeriosis were reported in England. We received<br />

a clinical questionnaire for 81 % <strong>of</strong> cases (N=1141). Eighty-two percent<br />

(N=934) had one or more underlying medical conditions and we recorded 1239<br />

ICD-10 codes on co-morbidities from these 934 cases. The ≥ 60 years age group<br />

comprised 76 % <strong>of</strong> all cases and 77 % <strong>of</strong> all co-morbidities. Overall, the highest comorbidity<br />

rates were diseases <strong>of</strong> the liver (192.9 episodes per million [95 % CI:<br />

150.9-242.9]), systemic connective tissue disorders (163.1 [108.4-235.7]), malignancies<br />

<strong>of</strong> the lymphoid and haema<strong>to</strong>poietic tissue (137.5 [118.5-158.7]), alcoholism<br />

(107.5 [80.8-140.3]), renal failure (103.5 [83.3-127.3]), diabetes (98.4 [76.8-124.1]),<br />

hypertensive disease (71.7 [44.9-108.5]) and malignancies <strong>of</strong> the eye, brain & other<br />

parts <strong>of</strong> CNS (66.3 [35.3-113.3]). We have highlighted several underlying conditions<br />

not previously thought <strong>to</strong> be strongly associated with listeriosis. The extent<br />

<strong>to</strong> which these co-morbidities are correlated with each other requires further investigation<br />

<strong>to</strong> enable much better, targeted prevention.<br />

Risk fac<strong>to</strong>rs for death in Listeria monocy<strong>to</strong>genes infection, England and<br />

Wales, 1990 <strong>to</strong> 2008<br />

Mook, P., Grant, K. and Gillespie, I.<br />

Health Protection Agency, UK<br />

Listeriosis, caused by the Gram positive bacterium Listeria monocy<strong>to</strong>genes, is a<br />

rare but severe food borne disease. The elderly, pregnant and the seriously ill are<br />

most <strong>of</strong>ten affected, with high mortality rates in all patient groups. The epidemiology<br />

<strong>of</strong> listeriosis in England & Wales has changed in recent years, with an average<br />

<strong>of</strong> 179 cases reported annually from 2000 <strong>to</strong> 2008 compared with 110 cases on average<br />

between 1990 and 1999. Much <strong>of</strong> the increase has occurred in patients > 59<br />

years presenting with bacteraemia but without central nervous system involvement.<br />

To examine fac<strong>to</strong>rs influencing L. monocy<strong>to</strong>genes mortality, non-pregnancy<br />

associated cases resident in England & Wales reported <strong>to</strong> national surveillance<br />

between 1990 and 2008 (N = 1832) were analysed further. The overall mortality<br />

rate for this period was 39 % (N = 719). Univariable and subsequent multivariable<br />

analyses were performed and the role <strong>of</strong> key fac<strong>to</strong>rs (season, age, underlying conditions<br />

and treatment) in this outcome investigated. Our initial findings suggest<br />

that, for a disease which disproportionately affects the elderly and the infirm, mortality<br />

is greatest in older patients and in those with underlying conditions. Underlying<br />

conditions themselves (particularly alcoholism and malignancies <strong>of</strong> the<br />

lymphatic/haema<strong>to</strong>poietic system and the breast) appear <strong>to</strong> be more important<br />

than the treatment they were receiving for an underlying condition, even when<br />

the effect <strong>of</strong> age is considered. Given that listeriosis is largely a food-borne disease,<br />

and t<strong>here</strong>fore preventable, it would seem appropriate <strong>to</strong> actively target high<br />

risk groups with specific advice on what foods <strong>to</strong> avoid in order <strong>to</strong> minimise the<br />

risk <strong>of</strong> listeriosis.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

A discrete event model <strong>to</strong> track Listeria monocy<strong>to</strong>genes<br />

in the retail environment<br />

Pouillot, R. 1 and Gallagher, D. 2<br />

1. CFSAN/FDA, USA<br />

2. Va Tech, USA<br />

A recent comparative Listeria monocy<strong>to</strong>genes (Lm) risk assessment predicts that <strong>of</strong><br />

the listeriosis cases attributed <strong>to</strong> deli meat, a majority are associated with deli<br />

meat sliced and packaged at retail. T<strong>here</strong> is a need <strong>to</strong>: 1) identify potential sources<br />

and practices that contribute <strong>to</strong> Lm contamination <strong>of</strong> food at the retail level; and 2)<br />

identify retail practices that would reduce or eliminate Lm contamination <strong>of</strong><br />

ready-<strong>to</strong>-eat foods prepared and sold <strong>to</strong> consumers. Within the scope <strong>of</strong> an interagency<br />

U.S. Food and Drug Administration (FDA) and U.S. Department <strong>of</strong> Agriculture/Food<br />

Safety and Inspection Service (USDA/FSIS) risk assessment <strong>of</strong> Lm at<br />

retail, a dynamic discrete-event model that tracks Lm cells at various locations<br />

over time in a retail setting was developed. This model considers i) the transfer <strong>of</strong><br />

bacteria from some defined compartment <strong>to</strong> others (unsliced products, sliced<br />

products, food contact surfaces, non food contact surfaces, slicer, gloves, hands,<br />

etc.); ii) the bacterial inactivation through cleaning and disinfection; and iii) the<br />

bacterial growth according <strong>to</strong> the physic and chemical environment (temperature,<br />

pH, water activity, presence <strong>of</strong> growth inhibi<strong>to</strong>rs). Sequences <strong>of</strong> events were derived<br />

using specifically acquired data from an observational study <strong>of</strong> food handling<br />

practices in retail deli departments. Additional data on the transmission <strong>of</strong><br />

Lm in the retail environment is being collected through labora<strong>to</strong>ry studies. Moreover,<br />

this risk assessment is designed <strong>to</strong> evaluate the relative effectiveness <strong>of</strong> retail<br />

Lm control measures along with other risk management scenarios. The major uncertainty<br />

is in the potential existence <strong>of</strong> niches and their interaction with the retail<br />

environment. This presentation will describe this model, provide an illustration<br />

using current data and highlight the data gaps that mainly influence the outputs.<br />

Clinical and epidemiological aspects <strong>of</strong> listeriosis in Israel<br />

Hershko-Klement, A. 1, Eliav, H. 2, Valinsky, L. 3, Schechner, V. 4, Braun, E. 5, Paitan, Y. 1,<br />

Block, C. S. 2 and Nir-Paz, R. 2<br />

1. Meir medical center, Kfar-Saba, Israel<br />

2. Hadassah Hebrew University Medical Center, Israel<br />

3. Israel Ministry <strong>of</strong> Health, Central Labora<strong>to</strong>ries<br />

4. Tel-Aviv Souraski medical Center, Tel Aviv, Israel<br />

5. Rambam medical center, Haifa, Israel<br />

Listeria monocy<strong>to</strong>genes (LM) is a ubiqui<strong>to</strong>us foodborne pathogen. Invasive infections<br />

are rare, mainly affecting elderly people, the immunocompromised and<br />

neonates. In pregnancy it causes fetal loss, preterm delivery and neonatal morbidity.<br />

Although the incidence <strong>of</strong> listeriosis is low compared with other enteropathogens,<br />

it carries a high mortality. Our aim was <strong>to</strong> characterize LM morbidity,<br />

risk fac<strong>to</strong>rs and molecular epidemiology in Israel <strong>to</strong> improve understanding<br />

<strong>of</strong> the disease burden in Israel and assist public health policy-making. We performed<br />

a nationwide retrospective study <strong>of</strong> all LM cases during 1998 – 2007. Cases<br />

were actively sought in the records <strong>of</strong> all hospital-based clinical microbiology labora<strong>to</strong>ries,<br />

the national reference LM labora<strong>to</strong>ry and LM cases reported <strong>to</strong> the district<br />

physicians. 481 cases were identified, <strong>of</strong> which 166 were pregnancy-associated.<br />

The yearly incidence peaked in 2006 and 2007 at 9.6 cases/million, which is<br />

almost 5 times than the current rate in the USA. The incidence in pregnancy varied<br />

between 5 and 25 cases/100,000 pregnancies/year. Annual rates above 15<br />

cases/100,000 occurred several times. The perinatal case fatality rate was 47 %,<br />

comprising a fetal loss <strong>of</strong> 38.2 % and additional 7.9 % neonatal mortality. Fetal survival<br />

improved as pregnancy age advanced (P < 0.05). LM associated late abortion<br />

occurred in 26.6 % and preterm labor in 46.8 %. A single maternal death was<br />

recorded. The case fatality in non-pregnancy cases was 18 %. Additionally a gradual<br />

increase in elderly occurred during the years and reached 42.7 cases/million/year.<br />

Geospatial analysis revealed 2 districts with a higher incidence <strong>of</strong> listeriosis. Molecular<br />

analysis (PFGE) <strong>of</strong> LM isolates suggested that a third <strong>of</strong> LM pregnancy associated<br />

cases were caused by a single clone. In Israel, Listeria monocy<strong>to</strong>genes is an<br />

important foodborne pathogen causing appreciable morbidity and mortality. Further<br />

studies <strong>of</strong> the sources, molecular epidemiology and specific virulence determinants<br />

should be undertaken.<br />

R E F E R E N C E<br />

C / O<br />

28<br />

R E F E R E N C E<br />

C / O<br />

29<br />

47


R E F E R E N C E<br />

C / O<br />

30<br />

R E F E R E N C E<br />

C / O<br />

31<br />

48<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Human isolates <strong>of</strong> Listeria monocy<strong>to</strong>genes during<br />

half a century in Sweden<br />

Lopez-Valladares, G., Danielsson-Tham, M.-L., Vishal Singh, P. and Tham, W.<br />

School <strong>of</strong> Hospitality, Culinary Arts & Meal Sciences, Örebro University, Grythyttan, Sweden<br />

Over 800 isolates <strong>of</strong> Listeria monocy<strong>to</strong>genes have been collected from cases <strong>of</strong> invasive<br />

listeriosis in Sweden – the first from 1958 and the latest from spring 2010.<br />

All isolates are serotyped and characterized with pulsed-field gel electrophoresis<br />

(PFGE) and Asc I restriction enzyme. From 1972 <strong>to</strong> 1995, serovar 4b was the predominant<br />

serovar in human invasive listeriosis in Sweden. During the 1980s, some<br />

listeriosis outbreaks caused by L. monocy<strong>to</strong>genes serovar 4b, belonging <strong>to</strong> a special<br />

PFGE type, were diagnosed in the USA and Europe and were due <strong>to</strong> consumption<br />

<strong>of</strong> dairy products. The predominant strain in listeriosis cases in Sweden during<br />

the same time belonged <strong>to</strong> this type. The hypothesis is that a majority <strong>of</strong> human<br />

listeriosis serovar 4b strains in Sweden came from s<strong>of</strong>t cheeses imported from<br />

Mediterranean countries. The hunt for L. monocy<strong>to</strong>genes serovar 4b in food production<br />

plants is intensive in EU and USA and during recent years, cheeses have<br />

had higher microbiological quality due <strong>to</strong> certified dairy farms and increased microbiological<br />

control. This may be the reason for the decrease <strong>of</strong> serovar 4b cases<br />

<strong>of</strong> listeriosis in Sweden. In 1996, serovar 1/2a became the major serovar in listeriosis<br />

cases in Sweden. The consumption <strong>of</strong> gravad and cold-smoked salmons has<br />

increased in Sweden, and in a recent study, 12.9 % <strong>of</strong> ready-<strong>to</strong>-eat vacuum-packed<br />

gravad salmon and 28.0 % <strong>of</strong> cold-smoked salmons harboured L. monocy<strong>to</strong>genes.<br />

The maximum number <strong>of</strong> L. monocy<strong>to</strong>genes was 1500 cfu/g product. The most<br />

common PFGE types found in human cases <strong>of</strong> listeriosis in Sweden <strong>to</strong>day are also<br />

the types frequently encountered in vacuum-packed cold-smoked and gravad<br />

salmon/rainbow trout.<br />

Estimated incubation periods for listeriosis vary according <strong>to</strong> clinical<br />

form <strong>of</strong> disease<br />

Goulet, V., King, L., Vaillant, V. and De Valk, H.<br />

Institut de Veille Sanitaire, France<br />

Data on the incubation period <strong>of</strong> listeriosis are scarce. The incubation period can<br />

be calculated precisely when a patient has had a single exposure <strong>to</strong> a confirmed<br />

source <strong>of</strong> contamination. This information can be collected during the investigation<br />

<strong>of</strong> point-source food borne outbreaks. Our study aimed <strong>to</strong> estimate the listeriosis<br />

incubation period using available outbreak investigation data. Cases <strong>of</strong> invasive<br />

listeriosis from confirmed food borne point-source listeriosis outbreaks<br />

with precisely documented incubation periods and clinical forms <strong>of</strong> infection were<br />

identified during the period 1985 – 2008 by literature review and from non-published<br />

French outbreaks. Data from 9 published papers and from 6 French outbreaks<br />

were analysed. A precise incubation period was documented for 28 cases: 13<br />

with central nervous system (CNS) involvement (10 French outbreak and 3 published<br />

cases), 15 pregnancy-associated cases (12 French outbreak and 3 published<br />

cases). A confirmed food vehicle was identified for all outbreaks (rice salad, icecream,<br />

cheese, meat-based products). The overall median incubation period was 17<br />

days (range: 2 – 88 days). A longer incubation period was observed for pregnancyassociated<br />

cases (median: 28 days (range: 14 – 88 days)) than for cases with CNS involvement<br />

(median: 10 days (range: 2 – 19 days)). Insufficient data did not allow<br />

for analysis <strong>of</strong> the incubation period <strong>of</strong> the bacteraemic form <strong>of</strong> infection. Our results<br />

suggest that the incubation period for listeriosis varies according <strong>to</strong> the clinical<br />

form <strong>of</strong> disease. A longer incubation period was observed for pregnancy-associated<br />

cases than for cases with CNS involvement. This information could have<br />

implications for the investigation <strong>of</strong> food borne listeriosis outbreaks as the incubation<br />

period is used <strong>to</strong> determine the time period for which a food his<strong>to</strong>ry is collected.<br />

Adapting this time period according <strong>to</strong> the clinical form <strong>of</strong> infection could<br />

facilitate a more precise identification <strong>of</strong> food products likely <strong>to</strong> be the source <strong>of</strong><br />

contamination.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Listeria monocy<strong>to</strong>genes in food and animals<br />

in the European Union in 2008<br />

da Silva Felício, M. T., Rizzi, V., Boelaert, F. and Makela, P.<br />

Scientific Cooperation and Assistance Department, EFSA, Italy<br />

The Directive 2003/99/EC on Zoonoses obligates the European Union (EU) Member<br />

States (MSs) <strong>to</strong> collect data <strong>of</strong> zoonoses, zoonotic agents, antimicrobial resistance<br />

and food-borne outbreaks. These data are reported <strong>to</strong> the European Food<br />

Safety Authority (EFSA) who publishes the Community Summary Report. In 2008,<br />

a large number <strong>of</strong> investigations (128,000) concerning ready-<strong>to</strong>-eat (RTE) foodstuffs<br />

were reported by 26 MSs. The food categories most <strong>of</strong>ten covered were RTE<br />

meat products, cheeses and fishery products. L. monocy<strong>to</strong>genes was seldom detected<br />

above the legal safety limit <strong>of</strong> 100 cfu/g from RTE foods and findings over<br />

this limit were most <strong>of</strong>ten reported from fishery products, cheeses, meat products<br />

and sandwiches at levels <strong>of</strong> 0.2 – 0.5 % in the EU. L. monocy<strong>to</strong>genes was isolated<br />

from both cheeses made from raw or low-heat-treated milk and pasteurised milk<br />

as well as from s<strong>of</strong>t/semi-s<strong>of</strong>t and hard cheeses. L. monocy<strong>to</strong>genes was most <strong>of</strong>ten<br />

detected in s<strong>of</strong>t and semi-s<strong>of</strong>t cheeses made from pasteurised milk. L. monocy<strong>to</strong>genes<br />

was also reported, generally at a relatively low prevalence, from various animal<br />

species in 2008, demonstrating that animals act as one reservoir <strong>of</strong> Listeria<br />

bacteria although they rarely serve as a direct source <strong>of</strong> human infections. The<br />

highest prevalence <strong>of</strong> L. monocy<strong>to</strong>genes was found in sheep, goats and cattle. Reported<br />

data on the findings in RTE foods may be used <strong>to</strong> guide food controls carried<br />

in MSs <strong>to</strong> ensure compliance with L. monocy<strong>to</strong>genes criteria. Compared <strong>to</strong><br />

previous years the quality <strong>of</strong> the data received improved as regards reporting <strong>of</strong> the<br />

stage <strong>of</strong> sampling and the use <strong>of</strong> appropriate test methods that has eased the assessment<br />

<strong>of</strong> compliance with Listeria criteria at Community level.<br />

Listeria monocy<strong>to</strong>genes infection in the over 60s in England<br />

between 2005 and 2008: a retrospective case-control study utilising<br />

market research panel data<br />

Gillespie, I. A., Mook, P., Little, C. L. and Grant, K.<br />

Health Protection Agency, UK<br />

Listeriosis is a rare but life-threatening foodborne disease and, t<strong>here</strong>fore, the investigation<br />

<strong>of</strong> fac<strong>to</strong>rs which might alter people’s risk <strong>of</strong> infection is important <strong>to</strong> inform<br />

on prevention and control. The incidence <strong>of</strong> listeriosis in England has doubled<br />

since 2001, with a largely unexplained increase in people aged ≥ 60 years<br />

presenting with bacteraemia without central nervous system involvement. Standardised<br />

epidemiological data has been sought on cases <strong>of</strong> listeriosis reported in<br />

England since 2005, but the value <strong>of</strong> the data accrued is limited without some<br />

knowledge <strong>of</strong> exposure prevalence in the population at risk <strong>of</strong> listeriosis. The exposures<br />

<strong>of</strong> listeriosis cases aged ≥ 60 years reported in England from 2005 – 2008<br />

were compared <strong>to</strong> those <strong>of</strong> market research panel members representing the same<br />

population and time period. Exposures were grouped <strong>to</strong> facilitate comparison.<br />

Odds ratios and 95 % confidence intervals were calculated. Cases were more likely<br />

than panel members <strong>to</strong> report cold cooked meats (beef and pork, but not chicken),<br />

smoked fish (specifically salmon) and shellfish (prawns), dairy products (milk and<br />

certain cheeses) and mixed salads. They were less likely <strong>to</strong> report the consumption<br />

<strong>of</strong> sandwiches and fresh vegetables. To our knowledge this is the first time that<br />

market research data has been applied <strong>to</strong> infectious disease epidemiology in this<br />

way. The diversity <strong>of</strong> high-risk food exposures reflects the ubiquity <strong>of</strong> the microorganism<br />

in the environment and/or the susceptibility <strong>of</strong> those at risk, and suggests<br />

that a wider variety <strong>of</strong> foods can give rise <strong>to</strong> listeriosis. Food safety advice on<br />

avoiding listeriosis should be adapted accordingly.<br />

R E F E R E N C E<br />

C / O<br />

32<br />

R E F E R E N C E<br />

C / O<br />

33<br />

49


R E F E R E N C E<br />

C / O<br />

34<br />

R E F E R E N C E<br />

C / O<br />

35<br />

50<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Better and faster typing, MLVA – shall we play <strong>to</strong>gether?<br />

Larsson, J. T. 1, Roussel, S. 2 and Moller Nielsen, E. 1<br />

1. Statens Serum Institut, Copenhagen, Denmark<br />

2. Agence Française de sécurité sanitaire des aliments, Maisons-Alfort, France<br />

In 2007 and 2008 t<strong>here</strong> were four papers published on the subject <strong>of</strong> MLVA (Multi<br />

Locus VNTR [Variable Number <strong>of</strong> Tandem Repeats] Analysis) in L. monocy<strong>to</strong>genes.<br />

During the same period a pro<strong>to</strong>col with an extensive bioinformatic approach was<br />

developed at SSI. The yet unpublished SSI MLVA pro<strong>to</strong>col was created after analysis<br />

<strong>of</strong> 20 genomes. Ten VNTRs were chosen on the basis <strong>of</strong> both nucleotide and<br />

amino acid sequences analysis. The VNTRs have repeat lengths ranging from 6 <strong>to</strong><br />

15 basepairs. Degenerate primers were designed <strong>to</strong> match the diversity in the<br />

genomes, taking good care not <strong>to</strong> include any gaps (which were present in several<br />

loci/genomes). Finally three multiplex PCR groups were created and the VNTRs<br />

were analysed by electrophoresis. The SSI method has been evaluated with the<br />

last years Danish human listeriosis cases and the North American ILSI Listeria<br />

collection. Results indicate a very close match <strong>to</strong> two-enzyme PFGE and an even<br />

better separation <strong>of</strong> isolates. These results in concert with the reduced time and<br />

cost for using MLVA for typing started <strong>of</strong>f an incentive for unification and standardisation<br />

<strong>of</strong> MLVA in L. monocy<strong>to</strong>genes. In a recently started cross labora<strong>to</strong>ry<br />

study, a comparison <strong>of</strong> the pro<strong>to</strong>col published by Kate Sperry (2007) and the pro<strong>to</strong>col<br />

developed at SSI is under way. The methods are evaluated <strong>to</strong>gether with traditional<br />

typing methods such as PFGE and agglutination serotyping. The MLVA<br />

pro<strong>to</strong>cols are tested using capillary and agarose gel electrophoresis. Strain collections<br />

includes; a selection <strong>of</strong> the WHO subtyping study from 1986 and a selection<br />

from food and human isolates samples from France and Denmark respectively.<br />

The chosen method might include a reduced set <strong>of</strong> loci and/or a mixture <strong>of</strong> the<br />

two assays. The results from this cooperation will lead <strong>to</strong> a recommendation on a<br />

MLVA pro<strong>to</strong>col for use in Denmark and France.<br />

comK prophage junction fragments in Listeria monocy<strong>to</strong>genes contain<br />

SNPs that differentiate subclones <strong>of</strong> ECII and ECIII that are unique <strong>to</strong><br />

individual meat and poultry processing plants in the U.S.<br />

Knabel, S. J.<br />

Department <strong>of</strong> Food Science, Penn State University, USA<br />

Many outbreaks <strong>of</strong> listeriosis in the U.S. are due <strong>to</strong> epidemic clones II and III, especially<br />

those associated with ready-<strong>to</strong>-eat meat and poultry products. Both <strong>of</strong><br />

these epidemic clones have backbone genomes that are highly conserved, but contain<br />

a hypervariable 40 Kb comK prophage. All isolates analyzed in the present<br />

study that were suspected <strong>of</strong> being ECII or ECIII were first confirmed as such by<br />

multiplex PCR and multi-virulence-locus sequence typing (Chen and Knabel,<br />

2007) and included those from the 1998 – 1999 hot dog and 2002 turkey deli outbreaks<br />

in the U.S., those from multiple states that were generated as part <strong>of</strong> USDA<br />

FSIS's meat and poultry plant surveillance program, and those that were previously<br />

isolated from two turkey processing plants in two different states (Eifert et<br />

al. 2005). Upstream and downstream prophage junction fragments in all ECII and<br />

ECIII isolates were amplified using 1/4 and 2/3 primer sets modified from Loessner<br />

et al. (2000) and subsequently sequenced. The junction fragments in ECIII<br />

isolates from a sporadic case in 1988 and from an outbreak in 2000 due <strong>to</strong> turkey<br />

deli meat in the U.S. were also amplified and sequenced. Cluster analysis based on<br />

junction fragment sequences revealed the presence <strong>of</strong> five subclones <strong>of</strong> ECII, the<br />

ECIII subclone and the Canadian 2008 subclone, which were unique <strong>to</strong> individual<br />

meat and poultry processing plants. However, the ECII subclone that caused the<br />

2002 outbreak was found in three different processing plants, two <strong>of</strong> which were<br />

originally associated with this outbreak. We speculate that evolution <strong>of</strong> these subclones<br />

may be due <strong>to</strong> extensive recombination and subsequent niche-specific<br />

adaptation acting on the comK prophage, which may represent a “Rapid Adaptation<br />

Island” in Listeria monocy<strong>to</strong>genes. Genes that were unique <strong>to</strong> the comK<br />

prophage were identified as putative “Adap<strong>to</strong>ns” and their possible roles in colonization,<br />

transmission between and within meat and poultry plants, and subsequent<br />

contamination <strong>of</strong> foods and humans will be discussed.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

ORAL PRESENTATIONS // O / 24–36<br />

Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes measured by multiple-locus<br />

variable-number tandem repeat analysis (MLVA)<br />

Hyytia-Trees, E., Sabol, A., Graves, L. and Ribot, E.<br />

Centers for Disease Control and Prevention, USA<br />

Listeria monocy<strong>to</strong>genes is a genetically and phenotypically diverse organism, encompassing<br />

at least three genetically distinct lineages that appear <strong>to</strong> differ in their<br />

likelihood <strong>of</strong> causing human and animal disease. Classification <strong>of</strong> isolates in<strong>to</strong><br />

these three lineages has been confirmed by different subtyping methods, and also<br />

correlates with serotype classification. In the present study, 250 epidemiologically<br />

unrelated isolates from 1981 <strong>to</strong> 2009 representing serotypes 4b (93), 1/2a<br />

(70), 1/2b (57), 3a (4), 4c (4), 4a (2), 1/2c (2) and untypeable (18) were characterized<br />

using multiple-locus variable-number tandem repeat analysis (MLVA). The<br />

isolate set also included one isolate from each <strong>of</strong> ten well characterized outbreaks.<br />

The data was analyzed in BioNumerics s<strong>of</strong>tware and a dendrogram was constructed<br />

by UPGMA clustering using the categorical coefficient. A minimal spanning<br />

tree was also created using the Manhattan coefficient. In the minimum spanning<br />

tree, a clonal complex was defined as a group <strong>of</strong> related patterns that differed<br />

from each other at a single locus by one or two repeats. A <strong>to</strong>tal <strong>of</strong> 139 unique patterns<br />

were detected among the 250 isolates. Thirty one MLVA patterns were associated<br />

with more than one isolate, with the most common pattern including 16<br />

isolates with isolation years varying from 1983 <strong>to</strong> 2007. With very few exceptions,<br />

the isolates clustered according <strong>to</strong> their lineage. Nineteen clonal complexes were<br />

detected, encompassing a <strong>to</strong>tal <strong>of</strong> 129 isolates. Two different serotypes (1/2a & 3a,<br />

1/2a & 1/2c) shared the same clonal complex in two instances. Five <strong>of</strong> the ten outbreak<br />

related isolates were located in three different clonal complexes that included<br />

nine or more isolates with multiple isolation years and origins in different<br />

continents. In conclusion, the MLVA scheme used in the present study is in accordance<br />

with the currently used lineage classification and indicated the presence<br />

<strong>of</strong> successful epidemic clones.<br />

R E F E R E N C E<br />

C / O<br />

36<br />

51


R E F E R E N C E<br />

D / O<br />

37<br />

R E F E R E N C E<br />

D / O<br />

38<br />

52<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 37– 46<br />

Time Temperature Indica<strong>to</strong>rs can be used as an effective Risk<br />

Management Tool for Listeria monocy<strong>to</strong>genes in Ready-To-Eat foods<br />

Koutsoumanis, K., Vaikousi, H. and Costas, B.<br />

Department <strong>of</strong> Food Science and Technology, School <strong>of</strong> Agriculture,<br />

Aris<strong>to</strong>tle University <strong>of</strong> Thessaloniki, Greece<br />

The objective <strong>of</strong> the present work was <strong>to</strong> evaluate the effectiveness <strong>of</strong> Time Temperature<br />

Indica<strong>to</strong>rs as risk management <strong>to</strong>ols for Listeria monocy<strong>to</strong>genes in RTE<br />

foods. In a previous study we presented a microbial Time Temperature Indica<strong>to</strong>r<br />

(TTI) system based on the growth and metabolic activity <strong>of</strong> a Lac<strong>to</strong>bacillus sakei<br />

strain. In the latter system, an irreversible color change <strong>of</strong> a chemical chromatic<br />

indica<strong>to</strong>r (from red <strong>to</strong> yellow) progressively occurs due <strong>to</strong> the pH decline as a result<br />

<strong>of</strong> L. sakei growth and metabolism in a selected medium. In this study we show<br />

that both the L. sakei growth and colour change <strong>of</strong> the microbial TTI system exhibit<br />

similar kinetic responses with the growth <strong>of</strong> L. monocy<strong>to</strong>genes. With an appropriate<br />

selection <strong>of</strong> the initial level <strong>of</strong> L. sakei in the system, the TTI end point<br />

(time at which a distinct visual color change <strong>to</strong> the final yellow was observed) can<br />

be adjusted in order <strong>to</strong> indicate a certain level <strong>of</strong> L. monocy<strong>to</strong>genes growth in the<br />

food during distribution and s<strong>to</strong>rage. Thus, the use <strong>of</strong> the microbial TTI can assure<br />

a maximum limit in the growth <strong>of</strong> the pathogen from production <strong>to</strong> consumption<br />

time by informing the consumers when this limit is exceeded in a product<br />

unit. The latter limit, which we call Growth Tolerance Criterion (GTC), can be<br />

considered as a performance criterion for the growth <strong>of</strong> L. monocy<strong>to</strong>genes during<br />

distribution and s<strong>to</strong>rage. The applicability <strong>of</strong> the microbial TTI in reducing consumer<br />

exposure <strong>to</strong> L. monocy<strong>to</strong>genes from the consumption <strong>of</strong> pate was evaluated<br />

in a simulation study. The TTI was appropriately adjusted <strong>to</strong> provide a GTC=3 logs<br />

CFU/g. The concentration <strong>of</strong> the pathogen and the colour <strong>of</strong> the TTI at the time <strong>of</strong><br />

consumption were estimated with a probabilistic approach using Monte Carlo<br />

simulation. The simulation results showed that the application <strong>of</strong> the TTI resulted<br />

in a significant reduction <strong>of</strong> consumer exposure <strong>to</strong> L. monocy<strong>to</strong>genes. Assuming<br />

that packages in which the TTI end point has been reached before consumption<br />

are discarded, the predicted percentage <strong>of</strong> consumed products contaminated with<br />

L. monocy<strong>to</strong>genes concentrations above 10 3 CFU/g was reduced from 5 % <strong>to</strong> 0.2 %<br />

with the use <strong>of</strong> TTI.<br />

Safety <strong>of</strong> salad leaves and herbs<br />

Garland, C. D. 1 and Clark, A. 2<br />

1. FWE Health, North Hobart, Tasmania, Australia<br />

2. Hous<strong>to</strong>ns Farm, Cambridge, Tasmania, Australia<br />

Comprehensive HACCP-based procedures have been implemented <strong>to</strong> prevent Listeria<br />

monocy<strong>to</strong>genes contamination <strong>of</strong> RTE (ready-<strong>to</strong>-eat) salad leaves and herbs in<br />

a commercial facility in Tasmania, Australia. Production has increased from 27<br />

<strong>to</strong>nnes in 1995 <strong>to</strong> 2000 <strong>to</strong>nnes in 2009, with cultivation now undertaken at 3 field<br />

sites and processing in a centralised fac<strong>to</strong>ry certified <strong>to</strong> third party audited food<br />

safety standards. The range <strong>of</strong> consumers is wide, including typical high-risk<br />

groups. More than 1200 samples <strong>of</strong> RTE products have been tested since 1995 and<br />

currently 300 RTE samples are tested at early and end <strong>of</strong> shelflife annually. To date,<br />

L. monocy<strong>to</strong>genes has been detected in one RTE sample only, an outstanding result.<br />

Control Points/Critical Control Points for minimising Listeria contamination in<br />

field conditions are implemented for: lettuce transplants; cultivation soil; catchment<br />

animals and wild life; irrigation water; harvesting; acceptance/rejection criteria<br />

for raw ingredients; potable water; transport <strong>of</strong> harvest <strong>to</strong> fac<strong>to</strong>ry. CPs/CCPs<br />

in the fac<strong>to</strong>ry relate <strong>to</strong>: restricted access; strict separation <strong>of</strong> work spaces in<strong>to</strong> lowmedium-high<br />

risk zones; temperature control; reduction <strong>of</strong> microbial loads; product<br />

packing; staff training; sensory moni<strong>to</strong>ring <strong>of</strong> product; physical and microbiological<br />

moni<strong>to</strong>ring. CPs/CCPs for transport <strong>to</strong> wholesaler and retailer include:<br />

sealed box packing, visual checking; sealed vehicle; temperature control/data logger;<br />

microbiological moni<strong>to</strong>ring. Data will be presented <strong>to</strong> illustrate key microbiological<br />

and physical CPs/CCPs relating <strong>to</strong>: animal access <strong>to</strong> the field, soil additives<br />

and composts; surfaces, equipment, potable water, irrigation water, raw<br />

ingredients and RTE products; fac<strong>to</strong>ry temperature control (ambient, water, product);<br />

removal <strong>of</strong> organisms by washing, centrifugation and infrared drying, and<br />

disinfection by chlorination (with au<strong>to</strong> adjustment <strong>of</strong> free available chlorine and<br />

pH); transport.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 37– 46<br />

Environmental fac<strong>to</strong>rs affect adhesion <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

<strong>to</strong> inert surfaces through flagellum expression<br />

Tresse, O.<br />

INRA, France<br />

Listeria monocy<strong>to</strong>genes could ad<strong>here</strong> <strong>to</strong> inert surfaces and form bi<strong>of</strong>ilms. Bi<strong>of</strong>ilms<br />

<strong>of</strong> L. monocy<strong>to</strong>genes constitute eventually a reservoir for cross- or re-contaminations<br />

<strong>of</strong> food products in food-processing plants. Bi<strong>of</strong>ilm initiation includes an irreversible<br />

attachment <strong>of</strong> cells (adhesion) <strong>to</strong> inert surfaces. Adhesion is t<strong>here</strong>fore<br />

crucial for the bi<strong>of</strong>ilm development and subsequently for dissemination and persistence<br />

<strong>of</strong> food pathogens. Adhesion assays from four <strong>to</strong> 101 L. monocy<strong>to</strong>genes<br />

strains from diverse origins (food, food-processing environment and clinical cases)<br />

were conducted in polystyrene and stainless steel 96-wells microtiter plates. Adhesion<br />

was also assessed after cell cultivation at different pHs, NaCl concentrations<br />

and temperatures. In parallel, flagella were observed after cultivation in the<br />

different environmental conditions using optical microscopy. Three naturally aflagellated<br />

strains, five environmental flaA mutant strains and one flagellum-paralyzed<br />

strain were also tested <strong>to</strong> verify the contribution <strong>of</strong> flagella <strong>to</strong> adhesion. Results<br />

indicated that attached cells were significantly higher at pHs greater or equal<br />

<strong>to</strong> 6, at 8 °C and 20 °C or at 0 % and 6 % NaCl than at pH 5, at 37 °C or at 11 % NaCl.<br />

Correlatively, flagella were not detected at pH 5, at 37 °C and at 11 % NaCl. This<br />

result indicates that the temperature is not the only environmental fac<strong>to</strong>r that<br />

could regulate flagellum expression in L. monocy<strong>to</strong>genes. NaCl concentrations and<br />

pHs closed <strong>to</strong> non-growing conditions inhibit flagellum expression. Naturally aflagellated<br />

strains and ∆flaA mutants ad<strong>here</strong>d at the same level <strong>of</strong> strains cultivated<br />

in flagellum inhibiting conditions. In addition, ∆flaA mutants ad<strong>here</strong>d equally<br />

while adhesion capability <strong>of</strong> the respective parental strains varied according <strong>to</strong><br />

the strain. Furthermore, adhesion was also affected in the flagellum-paralyzed<br />

strain indicating that motility plays an important role in adhesion capability. In<br />

conclusion, conditions <strong>of</strong> food conservation have an effect on adhesion capability<br />

<strong>of</strong> L. monocy<strong>to</strong>genes strains through flagellum regulation and flagella are responsible<br />

for adhesion variation among strains.<br />

Shelf-life labora<strong>to</strong>ry durability and challenge studies for<br />

Listeria monocy<strong>to</strong>genes in ready-<strong>to</strong>-eat foods: a presentation <strong>of</strong> the<br />

European technical guidance document intended for labora<strong>to</strong>ries<br />

Beaufort, A., Bergis, H., Cornue, M. and Lardeux, A.-L.<br />

Agence Française de Sécurité Sanitaire des Aliments (AFSSA), France<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen that may grow at refrigeration<br />

temperatures and may be present in ready-<strong>to</strong>-eat (RTE) foods. In annex I <strong>of</strong> regulation<br />

(EC) No 2073/2005 on microbiological criteria for foodstuffs, a specific attention<br />

<strong>to</strong> food safety criteria for L. monocy<strong>to</strong>genes in RTE foods is given and annex<br />

II <strong>of</strong> this regulation specifies that food business opera<strong>to</strong>rs (FBO’s) shall conduct, as<br />

necessary, studies <strong>to</strong> evaluate the growth <strong>of</strong> L. monocy<strong>to</strong>genes that may be present<br />

in the product during the shelf-life under reasonably foreseeable s<strong>to</strong>rage conditions.<br />

But this annex does not describe the procedure <strong>to</strong> conduct such shelf-life<br />

studies <strong>to</strong> ensure that the criterion <strong>of</strong> 100 L. monocy<strong>to</strong>genes/g is met over the entire<br />

intended shelf-life <strong>of</strong> the product. The European technical guidance document,<br />

intended for labora<strong>to</strong>ries conducting shelf-life studies for L. monocy<strong>to</strong>genes in RTE<br />

foods in collaboration with the FBO’s, provides recommendation on how <strong>to</strong> select,<br />

how <strong>to</strong> implement and how <strong>to</strong> perform the test(s) required. It describes the microbiological<br />

procedures for determining growth <strong>of</strong> L. monocy<strong>to</strong>genes using challenge<br />

tests and durability studies. Challenge tests, defined as tests providing information<br />

on the behaviour <strong>of</strong> L. monocy<strong>to</strong>genes artificially inoculated in the food,<br />

can be performed with two different objectives: (i)Classify whether a food does or<br />

does not support growth <strong>of</strong> L. monocy<strong>to</strong>genes; (ii) Quantify the growth <strong>of</strong> L. monocy<strong>to</strong>genes<br />

by assessing either the growth potential (δ) or the maximum growth rate<br />

(∝ max ). Durability studies allow an evaluation <strong>of</strong> the growth <strong>of</strong> L. monocy<strong>to</strong>genes in<br />

a naturally contaminated food during its s<strong>to</strong>rage according <strong>to</strong> reasonably foreseeable<br />

conditions. This technical document, prepared by the EU Community Reference<br />

Labora<strong>to</strong>ry for L. monocy<strong>to</strong>genes in collaboration with a working group consisting<br />

<strong>of</strong> 10 labora<strong>to</strong>ries, including nine National Reference Labora<strong>to</strong>ries for L.<br />

monocy<strong>to</strong>genes, is available since 15 December 2008 on the website <strong>of</strong> the European<br />

Commission, DG Health and Consumers.<br />

R E F E R E N C E<br />

D / O<br />

39<br />

R E F E R E N C E<br />

D / O<br />

40<br />

53


R E F E R E N C E<br />

D / O<br />

41<br />

R E F E R E N C E<br />

D / O<br />

42<br />

54<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 37– 46<br />

Successful strategies against Listeria monocy<strong>to</strong>genes in Switzerland<br />

Imh<strong>of</strong>, R.<br />

Agroscope Liebefeld-Posieux Research Station ALP, Switzerland<br />

In the aftermath <strong>of</strong> the massive Listeriosis outbreak in 1987 due <strong>to</strong> a s<strong>of</strong>t cheese<br />

made out <strong>of</strong> raw milk, the Swiss government decreed the creation <strong>of</strong> appropriate<br />

means <strong>to</strong> prevent a repetition <strong>of</strong> such a case. Agroscope Liebefeld-Posieux Research<br />

Station ALP was given the order <strong>to</strong> maintain a labora<strong>to</strong>ry for the detection<br />

<strong>of</strong> Listeria and <strong>to</strong> develop a nationwide Listeria moni<strong>to</strong>ring programme (LMP) in<br />

cooperation with the Swiss dairy industry. LMP covers about 70 <strong>to</strong> 80 % <strong>of</strong> the<br />

cheese production <strong>of</strong> Switzerland. The aims are <strong>to</strong> detect contamination in cheese<br />

production sites and ripening centres as early as possible, <strong>to</strong> s<strong>to</strong>p the spreading<br />

<strong>of</strong> Listeria by means <strong>of</strong> appropriate measures and above all <strong>to</strong> prevent contaminated<br />

products <strong>to</strong> be placed on the market. The programme is legally based on private<br />

law and guaranties confidentiality <strong>to</strong> its members. Although the results are<br />

not open <strong>to</strong> the public, ALP closely works <strong>to</strong>gether with the concerned public authorities<br />

in cases <strong>of</strong> enterprises with problems. ALP also created a special task<br />

force: By incidence <strong>of</strong> problems with Listeria the ALP consulting team can be demanded<br />

from any enterprise for analysis and advice as well as for direct participation<br />

in the process <strong>of</strong> complete redevelopments. Following defined proceedings<br />

the ALP team helps in cooperation with the enterprises own emergency team <strong>to</strong><br />

find individually designed solutions. Periodical controls in the following years<br />

pro<strong>of</strong> the success and the sustainability <strong>of</strong> interventions and worked out security<br />

concepts. The his<strong>to</strong>ry <strong>of</strong> outbreaks and the European RASFF notifications <strong>of</strong> the<br />

last 15 years concerning Switzerland manifest the success <strong>of</strong> this specific approach.<br />

Despite many ingenious efforts and product developments, t<strong>here</strong> exists – <strong>to</strong> our<br />

knowledge after 20 years <strong>of</strong> specific experience – no single measure or method <strong>to</strong><br />

solve a problem within the dairy food chain caused by L. monocy<strong>to</strong>genes. But observance<br />

<strong>of</strong> the principles <strong>of</strong> Good Manufacturing Practice, a specifically designed<br />

concept <strong>of</strong> security – which has <strong>to</strong> be lived by staff and personnel – besides <strong>to</strong> periodical<br />

controls guarantee that every production site is capable <strong>to</strong> produce and<br />

distribute safe and healthy dairy products.<br />

Absence <strong>of</strong> Listeria monocy<strong>to</strong>genes growth during raw milk<br />

cheesemaking: a modelling approach<br />

Jordan, K. 1, Schvartzman, S. 1,2, Butler, F. 2 and Tenenhaus-Aziza, F. 3<br />

1. Teagasc, Moorepark Food Research Centre, Ferrmoy, Cork, Ireland<br />

2. Biosystems Engineering, School <strong>of</strong> Agriculture, Food Science and Veterinary Medicine,<br />

University College Dublin, Ireland<br />

3. CNEIL, National Interpr<strong>of</strong>essionnal Center for Dairy Economy, Paris, France<br />

The presence <strong>of</strong> Listeria monocy<strong>to</strong>genes in certain foods and the risk that this poses<br />

<strong>to</strong> public health and food quality is still a problem. Currently, the field <strong>of</strong> food microbiology<br />

focuses on obtaining data on the behaviour <strong>of</strong> microorganisms in food,<br />

but the responses obtained provide little insight in<strong>to</strong> the relationship between<br />

physiological processes and growth or survival. This link can be made through<br />

mathematical models. In a simple form, a mathematical model is a simple mathematical<br />

description <strong>of</strong> a process. The application <strong>of</strong> mathematical models <strong>to</strong> food<br />

microbiology has been developed in recent years and now constitutes a new discipline<br />

named as Predictive Microbiology. However, most predictive models are<br />

based on labora<strong>to</strong>ry experiments in microbiological media under static conditions.<br />

As such models tend <strong>to</strong> be inaccurate, we have undertaken our experiments in a<br />

food system under dynamic conditions. Cheese was made with raw and pasteurised<br />

milk deliberately contaminated with L. monocy<strong>to</strong>genes. Listeria was moni<strong>to</strong>red<br />

through the manufacture and ripening period <strong>of</strong> the cheese. The results showed<br />

that L. monocy<strong>to</strong>genes did not grow during manufacture <strong>of</strong> raw milk cheese, but<br />

did grow during manufacture <strong>of</strong> pasteurised milk cheese. The data obtained for<br />

growth, survival and inactivation was modelled. The application <strong>of</strong> models that<br />

can explain the behaviour <strong>of</strong> Listeria in cheese and the further predictions that<br />

can be obtained from these models are useful for the improvement <strong>of</strong> ongoing research<br />

on biotraceability and for the better understanding <strong>of</strong> the general behaviour<br />

<strong>of</strong> these microorganisms under dynamic conditions, such as in dairy products.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 37– 46<br />

A predictive model <strong>to</strong> set high pressure processing criteria for<br />

Listeria monocy<strong>to</strong>genes inactivation on dry-cured ham<br />

Bover-Cid, S. 1, Belletti, N. 2, Garriga, M. 1 and Aymerich, T. 1<br />

1. IRTA. Food Technology, Spain<br />

2. Università di Bologna, Italy<br />

The aim <strong>of</strong> the work was <strong>to</strong> validate and use a predictive mathematical model<br />

for the inactivation <strong>of</strong> Listeria monocy<strong>to</strong>genes on dry cured ham by high hydrostatic<br />

pressure (HHP) processing, as a function <strong>of</strong> the technological parameters:<br />

intensity, length and fluid temperature. The model was previously developed<br />

from the inactivation data obtained after processing sliced dry-cured ham,<br />

spiked with L. monocy<strong>to</strong>genes, at different HHP conditions according <strong>to</strong> a central<br />

composite design: at 347 – 852 MPa; for 138 <strong>to</strong> 945 seconds; at 7.6 <strong>to</strong> 24.4 °C. The<br />

best fitting and most significant polynomial equation indicated that pressure<br />

and time were the most important fac<strong>to</strong>rs determining the inactivation extent.<br />

By contrast, temperature showed no significant influence within the assayed<br />

range. The statistically significant quadratic terms <strong>of</strong> pressure and time indicate<br />

the little effect observed below 450 MPa. Similarly, holding time longer than<br />

10 min did not result in a meaningful reduction <strong>of</strong> L. monocy<strong>to</strong>genes counts. The<br />

mathematical model was validated with data obtained from further experiments<br />

within the assayed experimental domain, including data from international scientific<br />

literature. The accuracy and bias fac<strong>to</strong>r were within the proposed acceptable<br />

values demonstrating the suitability <strong>of</strong> the model for predictive purposes.<br />

As a practical application <strong>of</strong> the validated model, the mathematical<br />

equation was used <strong>to</strong> predict the HHP process conditions needed <strong>to</strong> ensure the<br />

safety performance criteria established by different authors and organisations, in<br />

terms <strong>of</strong> logarithmic reductions <strong>of</strong> L. monocy<strong>to</strong>genes loads. The general performance<br />

criteria <strong>of</strong> 6 Log reduction, as equivalent <strong>to</strong> pasteurization, suggested<br />

by the Food and Drug Administration would be achieved with treatments above<br />

600 MPa, for instance applying 807 MPa for 5 min or 746 MPa for 10 min. More<br />

product-specific performance criteria proposed, i.e. 4D and 2.39D, would be<br />

achieved with 5 min treatments at 703 MPa and 613 MPa, respectively.<br />

Application <strong>of</strong> a validated predictive model <strong>to</strong> prevent growth<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes in ready-<strong>to</strong>-eat foods – importance<br />

for product development and risk management<br />

Mejlholm, O. and Dalgaard, P.<br />

Seafood & Predictive Microbiology, Technical University <strong>of</strong> Denmark<br />

Predictive microbiology is a most active research area within food microbiology<br />

and many successfully validated models are <strong>to</strong>day available and useful for assessment<br />

and management <strong>of</strong> microbial food safety and quality. Recently, an extensive<br />

growth and growth boundary model for Listeria monocy<strong>to</strong>genes was developed<br />

and successfully validated for ready-<strong>to</strong>-eat meat, seafood, poultry and non-fermented<br />

dairy products. This model is flexible and includes the effect <strong>of</strong> 12 environmental<br />

parameters (temperature, NaCl/a w , pH, smoke intensity, nitrite, CO 2 ,<br />

acetic acid, benzoic acid, citric acid, diacetate, lactic acid and sorbic acid) as well as<br />

their interactive effects. It can predict the growth boundary and t<strong>here</strong>by combinations<br />

<strong>of</strong> product characteristics and s<strong>to</strong>rage conditions that prevent growth <strong>of</strong> L.<br />

monocy<strong>to</strong>genes and this is most important in relation <strong>to</strong> EU regulations and codex<br />

guidelines. Nevertheless, growth boundary predictions alone are insufficient for<br />

formulation <strong>of</strong> safe foods because the likely variability in product characteristics<br />

and s<strong>to</strong>rage conditions must be taken in<strong>to</strong> account. For this, s<strong>to</strong>chastic models can<br />

be applied <strong>to</strong>gether with distributions <strong>of</strong> the relevant environmental conditions.<br />

Important parameters <strong>to</strong> control growth <strong>of</strong> L. monocy<strong>to</strong>genes in ready-<strong>to</strong>-eat foods,<br />

however, are not yet included in those models. We <strong>here</strong> present a simple alternative<br />

that allow variability in product characteristics and s<strong>to</strong>rage conditions <strong>to</strong> be<br />

taken in<strong>to</strong> account when conditions that prevent growth <strong>of</strong> L. monocy<strong>to</strong>genes are<br />

predicted. This approach relies on validated cardinal parameter growth models.<br />

The effect <strong>of</strong> interaction between environmental parameters is used <strong>to</strong> predict<br />

both the growth boundary and interface conditions within the no-growth region.<br />

Data from ready-<strong>to</strong>-eat foods will be used <strong>to</strong> show how these no-growth interfaces<br />

can be predicted in a sufficient distance from the growth boundary <strong>to</strong> account for<br />

variability in the environmental conditions.<br />

R E F E R E N C E<br />

D / O<br />

43<br />

R E F E R E N C E<br />

D / O<br />

44<br />

55


R E F E R E N C E<br />

D / O<br />

45<br />

R E F E R E N C E<br />

D / O<br />

46<br />

56<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

ORAL PRESENTATIONS // O / 37– 46<br />

Application <strong>of</strong> predictive microbiology <strong>to</strong> control the growth<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes – dairy products as an example<br />

Lobacz, A.<br />

Faculty <strong>of</strong> Food Sciences, University <strong>of</strong> Warmia and Mazury in Olsztyn, Poland<br />

Predictive microbiology is a relatively new scientific discipline which is used for<br />

modelling behaviour <strong>of</strong> microorganisms, mainly food borne pathogens, as a response<br />

<strong>to</strong> environment. It is based on assumption that the reaction <strong>of</strong> bacteria on particular<br />

environment fac<strong>to</strong>rs is reproducible, and on the basis <strong>of</strong> researches and observations<br />

made in past, it is possible <strong>to</strong> predict the microbiological behaviour in defined environment.<br />

Many mathematical models, with regard <strong>to</strong> Listeria monocy<strong>to</strong>genes, have<br />

been generated, which describe survival, inactivation or growth <strong>of</strong> this food borne<br />

pathogen as a response <strong>to</strong> environmental parameters. Models are generated usually<br />

in microbiological liquid media, although more <strong>of</strong>ten data are obtained in food products.<br />

In the following paper usefulness <strong>of</strong> predictive models describing the growth <strong>of</strong><br />

L. monocy<strong>to</strong>genes is presented. Application <strong>of</strong> tertiary models – ComBase Predic<strong>to</strong>r<br />

(www.combase.cc) and Pathogen Modeling Program (www.ars.usda.gov) is also discussed.<br />

Moreover, microbiological experiment was performed in order <strong>to</strong> evaluate<br />

behaviour <strong>of</strong> L. monocy<strong>to</strong>genes in Camembert type cheese during ripening and s<strong>to</strong>rage<br />

at following temperatures: 3, 6, 9, 12, 15 and 21 °C. Primary and secondary models<br />

were generated and their performance was checked by mathematical and graphical<br />

validation. Predictive microbiology is a new <strong>to</strong>ol <strong>to</strong> microbiological risk assessment.<br />

As a subdiscipline <strong>of</strong> food microbiology, it is an intensively developing field. Generated<br />

and validated predictive models become more accurate and more <strong>of</strong>ten used in<br />

food industry in order <strong>to</strong> eliminate food borne outbreaks.<br />

Characterization <strong>of</strong> Food Alert for Listeria monocy<strong>to</strong>genes in France in<br />

2008<br />

Leclercq, A. 1, Dusch, V. 2, Salah, S. 2, Laurent, E. 3, Chenal-Francisque, V. 1, Thierry-Bled, F. 4,<br />

Lecuit, M. 1, Goulet, V. 3 and Pihier, N. 2<br />

1. Institut Pasteur, NRC and WHO-CC for Listeria, Paris, France<br />

2. Direction Générale de l’Alimentation, Ministère de l’Alimentation, de l’Agriculture et de la Pêche, France<br />

3. Institut de Veille Sanitaire (InVS), France<br />

4. Direction Générale de la Concurrence, de la Consommation et de la Répression des Fraudes,<br />

Ministère de l’Economie, de l’Industrie et de l’Emploi, France<br />

Among measures established <strong>to</strong> assure a high level <strong>of</strong> protection <strong>of</strong> human health<br />

and consumers, the European regulation EC 178/2002 indicates that food business<br />

opera<strong>to</strong>rs shall notify all foodstuffs not in compliance with the food safety criteria<br />

for Listeria monocy<strong>to</strong>genes (Lm) described in regulation EC 2073/2005. A food alert<br />

is established when a foodstuff presenting a risk is on the market, or when rapid action<br />

(withdrawal or recall) is required. In the present study, our aim was <strong>to</strong> characterize<br />

all these French food alerts (FAs) for Lm in 2008. In case <strong>of</strong> FA, competent<br />

authorities asked the food business opera<strong>to</strong>rs or/and inspection services <strong>to</strong> send Lm<br />

food strains from own-checks and <strong>of</strong>ficial control <strong>to</strong> the NRC for Listeria <strong>to</strong> obtain<br />

their genoserotype and combined AscI/ApaI patterns by PFGE. Weekly, comparison<br />

<strong>of</strong> microbiological characteristics <strong>of</strong> food and human strains were performed on<br />

a period <strong>of</strong> 6 months and were sent <strong>to</strong> involved competent authorities and French<br />

Institute <strong>of</strong> Public Health Surveillance (InVS) with the aim <strong>of</strong> possible complementary<br />

investigation, if necessary. In 2008, 280 FAs for Lm have been registered, mainly<br />

at the retail level (184), from own-checks (215) and <strong>of</strong>ficial control (65). Among these<br />

280 FAs, Lm enumerations on sample(s) <strong>of</strong> FAs have been performed in 260 FAs<br />

[Range: 10-740,000 CFU/g]. Incriminated foods were mainly in decreasing order:<br />

meat products (134 FAs), raw milk cheeses (47 FAs) and seafood/fish products (40<br />

FAs). 179 FAs were only followed by food strains invoice [Range: 1-52 strains/FA; 100<br />

FAs with a single strain associated], representing 673 food strains. Genoserotypes <strong>of</strong><br />

food strains were at 55 % IIa, 21 % IVb, 19 % IIb and 4 % IIc. Combined AscI/ApaI<br />

patterns <strong>of</strong> strains varied from 1 <strong>to</strong> 27 per FA [137 showed single combined<br />

AscI/ApaI pattern]. 60 % <strong>of</strong> combined AscI/ApaI patterns for FAs food strains have<br />

been also observed for human strains in 2008. 84 FAs have food strains found microbiologically<br />

similar <strong>to</strong> human strains but no FA was linked with human cases in<br />

2008. Work is in progress for comparison <strong>of</strong> these data <strong>to</strong> 2007 and 2009 data. Even<br />

if high food contamination were sometimes observed, FAs were not at the origin or<br />

followed by human cases in 2008. In addition with the measures performed by the<br />

food business opera<strong>to</strong>rs in the context with their sanitary control plan, food management<br />

measures linked <strong>to</strong> FAs may have contributed in the French situation <strong>of</strong> no<br />

Lm outbreak since 2003.


AREA // E // Communication, risk perception and consumer practices – Social sciences in Listeria control<br />

ORAL PRESENTATIONS // O / 47– 49<br />

Efforts <strong>to</strong> update and perform health education on listeriosis<br />

in Los Angeles County, California, by the County <strong>of</strong> Los Angeles<br />

Department <strong>of</strong> Public Health<br />

Guevara, R. E.*<br />

County <strong>of</strong> Los Angeles Department <strong>of</strong> Public Health, USA<br />

Listeriosis became a relatively well-known disease in Los Angeles County after the<br />

large southern California listeriosis outbreak in 1985. However, 20 years later, the<br />

tasks <strong>of</strong> keeping the general population and medical community properly educated<br />

have not been easy with public health priorities shifting across <strong>to</strong>pics such as bioterrorism<br />

preparedness, MRSA, West Nile Virus, pandemic influenza, and most recently<br />

H1N1 influenza. In 2005, health education efforts on listeriosis were specific <strong>to</strong> only<br />

certain high-risk groups and needed a broader audience. Understanding the County<br />

<strong>of</strong> Los Angeles Department <strong>of</strong> Public Health’s (DPH) health education mission <strong>to</strong> encompass<br />

the general community, a multi-disciplinary team composed <strong>of</strong> epidemiologists,<br />

public health nurses, and health educa<strong>to</strong>rs produced new listeriosis education<br />

materials. Public health messages were developed by those who conducted case investigations<br />

by interviewing patients, relatives, and providers <strong>of</strong> listeriosis cases, and<br />

by those who performed epidemiologic surveillance. A standard <strong>of</strong> attractiveness was<br />

set before recruiting health educa<strong>to</strong>rs <strong>to</strong> help design the new education materials.<br />

Finding various images and pictures, health educa<strong>to</strong>rs drafted different designs for<br />

posters, flyers, and brochures. Public health nurses formed focus groups <strong>to</strong> help refine<br />

health education messages, particularly for materials translated in<strong>to</strong> Spanish. No<br />

systematic measurement for success <strong>of</strong> the health education materials was established;<br />

however, the materials continue <strong>to</strong> be used by DPH as various health care<br />

providers and organizations request them, and even when H1N1 influenza became<br />

the focus <strong>of</strong> public health in April 2009, listeriosis was the most common search term<br />

in the DPH website. This presentation describes the approach the DPH used <strong>to</strong> more<br />

effectively communicate the risks <strong>of</strong> listeriosis not just <strong>to</strong> those who are vulnerable <strong>to</strong><br />

the disease, but also <strong>to</strong> those who might know or even take care <strong>of</strong> them.<br />

* Participation Supported by IUFoST<br />

Awareness <strong>of</strong> listeriosis among Portuguese pregnant women<br />

Mateus, T. 1,3, Maia, R. L. 2 and Teixeira, P. 1<br />

1. CBQF / Escola Superior de Biotecnologia – Universidade Católica Portuguesa, Por<strong>to</strong>, Portugal<br />

2. CECLICO / Centro de Estudos Culturais, da Linguagem e do Comportamen<strong>to</strong> Faculdade de<br />

Ciências Humanas e Sociais – Universidade Fernando Pessoa, Por<strong>to</strong>, Portugal<br />

3. Escola Universitária Vasco da Gama, Coimbra, Portugal<br />

The role <strong>of</strong> Listeria monocy<strong>to</strong>genes as a foodborne pathogen was definitively recognized<br />

during the 1980s. This recognition was the consequence <strong>of</strong> a number <strong>of</strong> epidemic<br />

human outbreaks due <strong>to</strong> the consumption <strong>of</strong> contaminated foods, in Canada, the US<br />

and Europe. Listeriosis is especially severe in immunocompromised individuals such<br />

as pregnant women. The disease has a low incidence, although this is undeniably increasing,<br />

and a high fatality rate amongst those infected. In pregnant women listeriosis<br />

may cause miscarriage, foetal death or neonatal morbidity in the form <strong>of</strong> septicaemia<br />

and meningitis. Improved education concerning the disease, its transmission<br />

and prevention measures for immunocompromised individuals and pregnant women<br />

has been identified as a pressing need. The aim <strong>of</strong> this study was <strong>to</strong> evaluate the awareness,<br />

in terms <strong>of</strong> food safety and listeriosis, <strong>of</strong> pregnant women in the last trimester <strong>of</strong><br />

pregnancy or with babies up <strong>to</strong> 3 months old. For this purpose 956 women were questioned<br />

as <strong>to</strong> their knowledge and practices relating <strong>to</strong> what foods <strong>to</strong> avoid, good hygiene<br />

and cooking practices. Also <strong>of</strong> interest were the sources <strong>of</strong> information on Listeria<br />

spp. and listeriosis, and from whom and how women would like <strong>to</strong> receive such<br />

information. Two hundred and twenty nine women said they had not avoided any kind<br />

<strong>of</strong> food, whilst 306 said they had not made any particular changes <strong>to</strong> their food safety<br />

related habits when they prepared and cooked meals. Four hundred and seventy seven<br />

women considered that they had received enough information about nutrition during<br />

pregnancy, and that this information had been acquired from their doc<strong>to</strong>r, who was<br />

also the person considered the most competent <strong>to</strong> give this information. A hundred<br />

and seventeen women said they had heard about listeriosis, although only 58 <strong>of</strong> these<br />

had been able <strong>to</strong> describe listeriosis symp<strong>to</strong>ms. Women showed plenty <strong>of</strong> interest in<br />

new information about listeriosis, and they would like doc<strong>to</strong>rs <strong>to</strong> provide it, as well as<br />

by brochures and in the pregnancy advice book provided by the State Medical Service.<br />

It appears that plan is required, <strong>to</strong> raise awareness amongst health pr<strong>of</strong>essionals for<br />

the need for food safety education for pregnant women.<br />

R E F E R E N C E<br />

E / O<br />

47<br />

R E F E R E N C E<br />

E / O<br />

48<br />

57


R E F E R E N C E<br />

E / O<br />

49<br />

58<br />

AREA // E // Communication, risk perception and consumer practices – Social sciences in Listeria control<br />

ORAL PRESENTATIONS // O / 47– 49<br />

The development and progress <strong>of</strong> a risk-based strategy<br />

for the control <strong>of</strong> Listeria monocy<strong>to</strong>genes in New Zealand<br />

Castle, M. and Crerar, S.<br />

New Zealand Food Safety Authority<br />

The New Zealand Food Safety Authority (NZFSA) has developed a risk-based strategy<br />

for the control <strong>of</strong> Listeria monocy<strong>to</strong>genes. One <strong>of</strong> the objectives <strong>of</strong> this strategy<br />

is ‘no increase in the reported incidence <strong>of</strong> foodborne listeriosis over the five year<br />

period <strong>to</strong> 2013’. The strategy proposes a consistent and informed approach <strong>to</strong> managing<br />

the risk from L. monocy<strong>to</strong>genes, through: (i) the categorisation <strong>of</strong> ready-<strong>to</strong>eat<br />

foods in relation <strong>to</strong> their potential <strong>to</strong> support the growth <strong>of</strong> L. monocy<strong>to</strong>genes;<br />

(ii) the commissioning <strong>of</strong> scientific studies and the development <strong>of</strong> <strong>to</strong>ols <strong>to</strong> support<br />

industry and NZFSA make risk based decision; (iii) the application <strong>of</strong> management<br />

controls for industry using a risk-based approach; (iv) the implementation <strong>of</strong><br />

the principles in the Codex Guidelines on the Application <strong>of</strong> General Principles<br />

<strong>of</strong> Food Hygiene <strong>to</strong> the Control <strong>of</strong> Listeria monocy<strong>to</strong>genes; (v) risk communication<br />

for vulnerable consumers and food businesses. The strategy also recognises that<br />

t<strong>here</strong> is the potential for the increased incidence <strong>of</strong> listeriosis cases due <strong>to</strong> the impact<br />

<strong>of</strong> an aging population and the greater availability and range <strong>of</strong> chilled ready<strong>to</strong>-eat<br />

foods and the objective will t<strong>here</strong>fore only be achieved by working with the<br />

New Zealand food industry and consumers <strong>to</strong> enhance the control <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

The L. monocy<strong>to</strong>genes strategy was launched in 2008 and the initial response<br />

<strong>to</strong> the risk management <strong>to</strong>ols developed will be explored including uptake and response<br />

from industry.


POSTER PRESENTATIONS


R E F E R E N C E<br />

A / P<br />

00<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Compartmentalization <strong>of</strong> IFN-gamma and IL-6 recep<strong>to</strong>rs signalling<br />

in Listeria monocy<strong>to</strong>genes phagosomes: immune vesicles<br />

Ramos-Vivas, J., Carrasco-Marin, E., Madrazo-Toca, F., Fernandez-Prie<strong>to</strong>, L.,<br />

Rodriguez-Del Rio, E., Carranza-Cereceda, C. and Alvarez-Dominguez, C.<br />

Hospital Santa Cruz de Liencres and FMV-IFIMAV, Spain<br />

Cy<strong>to</strong>kine activated macrophages kill Listeria monocy<strong>to</strong>genes within the phagosomes.<br />

Here, we describe a novel listericidal Stat1 mediated signalling pathway involving<br />

IFN-γ and IL-6. This pathway is compartmentalized in<strong>to</strong> IFN-γ/IL-6<br />

phago-recep<strong>to</strong>somes, innate immunity vesicles that link LM destruction with specific<br />

immunity and IL-6 regulation. The Stat1-mediated signal triggered within<br />

this compartment, links Rab5a with the activation <strong>of</strong> the lysosomal enzyme effec<strong>to</strong>r,<br />

cathepsin-D, and with the induction <strong>of</strong> oxidative mechanisms such as the phox<br />

regula<strong>to</strong>rs, p67phox/Rac2-GTP, and the iNOS. Three downstream events link these<br />

listericidal IFN-γ/IL-6 phago-recep<strong>to</strong>somes with specific immunity: (i) a Stat1independent<br />

transformation in<strong>to</strong> MIIC-like vesicles regulated by Rab5a-GTP that<br />

allows the acquisition <strong>of</strong> αβ SDS-stable MHC-II dimers, cathepsin-D, Lamp-2 and<br />

Limp-2. (ii) A Stat1-dependent degradation <strong>of</strong> the Listeria immune-dominant antigen,<br />

listeriolysin O by cathepsin-D. (iii) A downstream Stat1 signal <strong>to</strong> regulate IL-<br />

6 production and drive a Th-1 cell-mediated immunity.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Antimicrobial susceptibility among Listeria monocy<strong>to</strong>genes isolates<br />

from non human sources in France over a ten year period<br />

Granier, S. A. 1, Moubarek, C. 2, Colaneri, C. 1, Roussel, S. 1, Courvalin, P. 2 and Brisabois, A. 1<br />

1. AFSSA, France<br />

2. Institut Pasteur, France<br />

Listeriosis is public health concern since 1980’s. This food-borne infectious disease<br />

has very high rates <strong>of</strong> hospitalization and fatality. It usually requires antimicrobial<br />

treatment. Recommendations are: Penicillin G or ampicillin combined or not with<br />

aminoglycosides. Antimicrobial resistance has increased among Gram positive bacteria<br />

during the last past years but very few data are available for Listeria monocy<strong>to</strong>genes,<br />

especially from non-human sources. In order <strong>to</strong> determine the frequency <strong>of</strong><br />

resistance within the species L. monocy<strong>to</strong>genes, a collection <strong>of</strong> 205 food and environmental<br />

isolates, randomly selected among a collection <strong>of</strong> 10 year period<br />

(1996 – 2006) was tested for antimicrobial susceptibility. This collection <strong>of</strong> non epidemiologically<br />

related isolates included major serotypes (1/2a,1/2b,1/2c and 4b<br />

strains) encountered, each <strong>of</strong> them presenting a unique PFGE pr<strong>of</strong>ile. The phenotype<br />

was determined using the microdilution method as recommended by CLSI.<br />

Most <strong>of</strong> the strains presented a wild type phenotype: susceptible <strong>to</strong> penicillin G,<br />

ampicillin, tetracycline, erythromycin, gentamicin, sulfonamides, trimethoprim,<br />

vancomycin and resistant <strong>to</strong> 3rd generation cephalosporines, quinolones. Only four<br />

strains displayed an original phenotype: two isolates (serotype 4b) were resistant <strong>to</strong><br />

erythromycin, two isolates (serotype 1/2a) were resistant <strong>to</strong> tetracycline, one <strong>of</strong> the<br />

two last isolate was additionally resistant <strong>to</strong> trimethoprim. PCR analysis detected<br />

the presence <strong>of</strong> erm(B) gene in the 2 erythromycin resistant strains, tet(M) in the<br />

tetracycline resistant strains and dfr(D) was identified in the unique trimethoprim<br />

resistant strain. Interestingly, int gene, involved in the mobility <strong>of</strong> conjugative<br />

transposons, was found in the strains harboring tet(M) gene. The present results do<br />

not indicate high frequency <strong>of</strong> resistance, nor large dissemination <strong>of</strong> resistance.<br />

These data should now be compared <strong>to</strong> the one obtained for human isolates. Nevertheless,<br />

those findings need <strong>to</strong> be regularly updated <strong>to</strong> make sure that the resistance<br />

situation in L. monocy<strong>to</strong>genes doesn’t shift.<br />

Five homologous small RNAs are involved in the response<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes <strong>to</strong> cell wall acting antibiotics<br />

Kiil Nielsen, P. and Kallipolitis, B.<br />

Department <strong>of</strong> Biochemistry and Molecular Biology, Denmark<br />

Small RNAs (sRNAs) are important regula<strong>to</strong>ry elements involved in a variety <strong>of</strong><br />

cellular responses, including stress <strong>to</strong>lerance and virulence. In Listeria monocy<strong>to</strong>genes<br />

five homologous sRNAs, named LhrC1-5, were identified recently by coimmunoprecipitation<br />

with the RNA chaperone protein Hfq. The objective <strong>of</strong> the<br />

present study is <strong>to</strong> improve our understanding <strong>of</strong> the role(s) <strong>of</strong> LhrC1-5 in L. monocy<strong>to</strong>genes.<br />

Our study involves the identification <strong>of</strong> fac<strong>to</strong>rs that influence the expression<br />

<strong>of</strong> LhrC1-5 and investigations <strong>of</strong> genes affected by LhrC1-5. A range <strong>of</strong><br />

different environmental conditions and regula<strong>to</strong>ry mutant strains were employed<br />

<strong>to</strong> identify fac<strong>to</strong>rs that affect the expression <strong>of</strong> LhrC1-5. The presence and stability<br />

<strong>of</strong> the RNA transcripts were determined by Northern blotting. Furthermore, expression<br />

<strong>of</strong> lhrC1-5 was followed by construction <strong>of</strong> promoter-lacZ fusions for all<br />

five sRNA-encoding genes. These studies revealed a differentiated transcription <strong>of</strong><br />

LhrC1-5 in response <strong>to</strong> the presence <strong>of</strong> cell wall acting antimicrobial agents, such<br />

as ethanol and β-lactam antibiotics. Cell wall acting agents have previously been<br />

shown <strong>to</strong> activate the two-component systems LisRK and CesRK in L. monocy<strong>to</strong>genes.<br />

Our experiments revealed that transcription <strong>of</strong> lhrC1-5 is completely dependent<br />

on the presence <strong>of</strong> a functional LisRK system, w<strong>here</strong>as the CesRK system<br />

appears <strong>to</strong> be important for a differentiated transcription <strong>of</strong> lhrC1-5 in<br />

response <strong>to</strong> the above-mentioned cell wall acting agents. Finally, a mutant strain<br />

lacking lhrC1-5 was constructed and characterized, and genes regulated by LhrC1-<br />

5 were investigated by comparing the global gene expression <strong>of</strong> a wild type strain<br />

vs. a ∆lhrC1-5 mutant strain. In summary, this study shows that L. monocy<strong>to</strong>genes<br />

responds <strong>to</strong> cell wall acting agents through a complex gene regula<strong>to</strong>ry network<br />

which includes five small non-coding RNAs.<br />

R E F E R E N C E<br />

A / P<br />

01<br />

R E F E R E N C E<br />

A / P<br />

02


R E F E R E N C E<br />

A / P<br />

03<br />

R E F E R E N C E<br />

A / P<br />

04<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Phenotypic analysis <strong>of</strong> selected listerial secretion mutants<br />

Halbedel, S.*, Galander, S. and Flieger, A.<br />

Robert Koch Institute, Germany<br />

To be secreted in<strong>to</strong> any extracellular compartment, proteins need <strong>to</strong> be translocated<br />

across the cy<strong>to</strong>plasmic membrane first. Protein translocation is an energy<br />

driven process and mediated by different sophisticated nano-machines that link<br />

energy consumption <strong>to</strong> substrate translocation. Bulk protein translocation in the<br />

Firmicutes is mediated by the Sec translocon <strong>to</strong>gether with the YidC proteins that<br />

have their function in the lateral release <strong>of</strong> translocation substrates in<strong>to</strong> the phospholipid<br />

bilayer. Listeria monocy<strong>to</strong>genes YidC homologues are encoded by the spoI-<br />

IIJ (lmo2854) and yqjG (lmo1379) genes. In addition, several minor secretion systems<br />

have been described that transport certain subsets <strong>of</strong> translocation<br />

substrates only, such as the Tat system, or that are supposed <strong>to</strong> be involved in protein<br />

secretion since they seem <strong>to</strong> form proteinaceous membrane pores. Among<br />

the latter is the Bacillus subtilis SpoIIIAH protein that recently was shown <strong>to</strong> be<br />

the mother cell-specific part <strong>of</strong> a channel connecting the mother cell with the prespore<br />

compartment in sporulating B. subtilis cells. A homologue <strong>of</strong> this gene can be<br />

found in L. monocy<strong>to</strong>genes as well (lmo0791), w<strong>here</strong>as the prespore-specific component<br />

(spoIIQ) and the other seven genes <strong>of</strong> the B. subtilis spoIIIA operon seem<br />

<strong>to</strong> be absent from the listerial chromosome. L. monocy<strong>to</strong>genes SpoIIIAH consists <strong>of</strong><br />

an N-terminal transmembrane helix and a putative extracellular domain at the Cterminus<br />

that shares similarity with the ring forming proteins <strong>of</strong> the YscJ/FliF<br />

family associated with type III protein secretion. We t<strong>here</strong>fore hypothesized that<br />

SpoIIIAH might be involved in secretion <strong>of</strong> proteins or other biomolecules in<strong>to</strong><br />

the extracellular space in L. monocy<strong>to</strong>genes. To test this hypothesis and <strong>to</strong> describe<br />

the impact <strong>of</strong> YidC homologues on protein secretion we generated strains with<br />

deletions <strong>of</strong> the spoIIIAH, yidC or yqjG genes. Here, we will present data describing<br />

the phenotypic and proteomic analysis <strong>of</strong> these mutants.<br />

* Participation Supported by IUFoST<br />

Distribution <strong>of</strong> serotypes and pulsotypes <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in pig farms (France 2008)<br />

Boscher, E.<br />

AFSSA, France<br />

This work was undertaken <strong>to</strong> estimate the prevalence <strong>of</strong> L. monocy<strong>to</strong>genes and, distribution<br />

<strong>of</strong> serotypes and genotypes in French farrow-<strong>to</strong>-finish pig farms in 2008,<br />

at the breeding pig level. A <strong>to</strong>tal <strong>of</strong> 730 faeces (10 per farm) were sampled from<br />

sows in 73 pig farms localized in Brittany, France. Detection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

was carried out according <strong>to</strong> the ISO 11290-1/A1 method. Isolates were serotyped<br />

and typed by PFGE. We found that 46.6 % <strong>of</strong> the farms (34/73) had at least one positive<br />

sample and 10.7 % <strong>of</strong> the samples (78/730) were positive for L. monocy<strong>to</strong>genes.<br />

A <strong>to</strong>tal <strong>of</strong> 125 strains were collected. Isolates belonged <strong>to</strong> four serotypes (1/2a, 1/2b,<br />

4b and 1/2c) with a percentage <strong>of</strong> 41.6 %, 36.0 %, 20.8 % and 1.6 % respectively. Out <strong>of</strong><br />

the 34 positive farms, 19, 12 and 3 had respectively 1, 2, and 3 serotypes. Moreover,<br />

21, 17, 11 and 1 positive farms had respectively the serotype 1/2a, 1/2b, 4b and 1/2c.<br />

The genetic diversity <strong>of</strong> the isolates collected in this study was very important<br />

(DI=0.986). In <strong>to</strong>tal, 50 genotypes were obtained; 25 for 1/2a, 15 for 1/2b, 9 for 4b<br />

and 1 for 1/2c. Moreover 17, 6, 4, 3, 3, and 1 positive farms had respectively 1, 2, 3, 4,<br />

5 and 6 genotypes. Furthermore, strains showing similar genotypes occurred in<br />

several farms; as example, genotypes G18, G12 and G10 were found in 4, 5 and 5<br />

farms respectively. This study provided recent valuable information on the occurrence<br />

<strong>of</strong> L. monocy<strong>to</strong>genes in French pig farms. This microorganism is prevalent in<br />

the sows in farrow-<strong>to</strong>-finish pig farms (46.6 %). Our work highlighted for some<br />

farms several serotypes and genotypes suggesting several sources <strong>of</strong> contamination.<br />

Furthermore, the serotypes found in this study are identical <strong>to</strong> those usually<br />

involved in human listeriosis in Europe (Goulet et al. 2008).


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Comparative phylogenomics <strong>of</strong> Listeria monocy<strong>to</strong>genes reveals<br />

an adaptation pr<strong>of</strong>ile<br />

Silveira Nalério, E. 1, Padilha Silva, W. 1, Stabler, R. 2 and Wren, B. W. 2<br />

1. Universidade Federal de Pelotas, Brazil<br />

2. London School <strong>of</strong> Hygiene and Tropical Medicine, UK<br />

Listeria monocy<strong>to</strong>genes is the causative agent <strong>of</strong> listeriosis which may cause a range<br />

<strong>of</strong> diseases from gastroenteritis <strong>to</strong> death. In fact, disease outcome can be related <strong>to</strong><br />

strain serotype/lineage thus molecular analyses has demonstrated that L. monocy<strong>to</strong>genes<br />

is a highly diverse species which can be grouped in<strong>to</strong> three lineages.<br />

Whole-genome microarray can be employed <strong>to</strong> study phylogenetic relationships<br />

among Listeria strains either species or serotype level, in addition <strong>to</strong> demonstrate<br />

differences on their virulence potential and/or environmental adaptation. The<br />

aim <strong>of</strong> this study was the whole genome comparison <strong>of</strong> L. monocy<strong>to</strong>genes strains<br />

from different origins. Ninety-nine L. monocy<strong>to</strong>genes strains from different geographical<br />

origins (Brazil, Denmark, Austria, Ireland, USA and unknown), including<br />

clinical strains (humans and animals), food and food industries strains were analyzed.<br />

DNA from all strains were competitively hybridized on a L. monocy<strong>to</strong>genes<br />

DNA microarray based on the whole-genome sequence L. monocy<strong>to</strong>genes EGD-e.<br />

DNA labeling and hybridization pro<strong>to</strong>col were followed according <strong>to</strong> Dorrell et al.<br />

(2001). Data acquisition, processing and comparative phylogenomics were performed<br />

as previously described by Stabler et al. (2006). Comparative phylogenomics<br />

clustered the L. monocy<strong>to</strong>genes strains in<strong>to</strong> two central clades which is<br />

representative <strong>of</strong> the two main lineages <strong>of</strong> this species. In addition each <strong>of</strong> these<br />

clades was divided in<strong>to</strong> two further subclades. Clade formation was independent <strong>of</strong><br />

the geographical origin <strong>of</strong> strains with the exception <strong>of</strong> the clade containing persistent<br />

strains (strains that persist in food-processing environment), w<strong>here</strong> none<br />

<strong>of</strong> the Brazilian strains were present. It was found 18 specific genes for lineage I<br />

strains (1/2a and 1/2c serotypes). These genes are related <strong>to</strong> carbohydrate metabolism,<br />

two component regula<strong>to</strong>ry system, ABC transporter complex and bvrB and<br />

bvrC genes. Significantly all persistent strains clustered <strong>to</strong>gether in the same lineage<br />

I clade. We achieved a set <strong>of</strong> unique genes belonging exclusively <strong>to</strong> L. monocy<strong>to</strong>genes<br />

persistent strains pointing <strong>to</strong> be responsible for its adaptation pr<strong>of</strong>ile.<br />

The genes are involved in stress resistance and are related <strong>to</strong> carbohydrate transport<br />

and metabolism, environmental information processing, signal transduction<br />

mechanisms, cell surface protein, amino acid transport and metabolism, nucleotide<br />

transport and metabolism, translation, cell wall biogenesis, replication, recombination<br />

and repair, transport <strong>of</strong> small molecules similar <strong>to</strong> ABC transporter,<br />

metabolism <strong>of</strong> lipids and unknown function. Interestingly from 14 virulence listed<br />

genes most <strong>of</strong> them were present in all studied L. monocy<strong>to</strong>genes strains with exception<br />

<strong>of</strong> inlE and inlG genes. These findings indicate that genetic variability <strong>of</strong> L.<br />

monocy<strong>to</strong>genes strains point <strong>to</strong> niche adaptation instead virulence differentiation<br />

despite <strong>of</strong> different origins. Persistent strains clustered suggesting genetic origin <strong>to</strong><br />

survival in this environment.<br />

Serotyping and PFGE patterns <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

from poultry meat<br />

Vasantrao Kurkure, N. 1, Kalorey, D. R. 1, Rodrigues, J. 2, Gunjal, P. 1 and Barbuddhe, S. B. 2<br />

1. Nagpur Veterinary College, Maharashtra Animal & Fishery Sciences University, India<br />

2. ICAR Research Complex for Goa, India<br />

Listeria monocy<strong>to</strong>genes is a gram positive, facultative, intracellular bacteria which is<br />

ubiqui<strong>to</strong>us in nature. Poultry meat and products frequently have been implicated in<br />

outbreaks <strong>of</strong> food borne illness. In India t<strong>here</strong> is practice <strong>to</strong> sell freshly slaughtered<br />

poultry meat. A <strong>to</strong>tal <strong>of</strong> 119 poultry meat samples were collected from poultry meat<br />

shops. These samples were processed for isolation <strong>of</strong> Listeria spp. by adopting standard<br />

pro<strong>to</strong>col. Out <strong>of</strong> 47 Listeria spp. recovered 17 isolates were confirmed as L.<br />

monocy<strong>to</strong>genes by biochemical and in vitro pathogenicity test. The L. monocy<strong>to</strong>gens<br />

isolates were subjected <strong>to</strong> multiplex PCR based serotyping and PFGE analysis. All<br />

the isolates were grouped as 4b, 4d and 4e serotypes. By PFGE analysis four Asc I<br />

pr<strong>of</strong>iles were observed. Thirteen isolates were observed <strong>to</strong> be <strong>of</strong> similar pr<strong>of</strong>ile indicating<br />

homogenecity among the isolates recovered from poultry meat. Present<br />

findings indicated the need <strong>of</strong> improvement in hygienic practices while slaughtering<br />

and sell <strong>of</strong> poultry meat as L. monocy<strong>to</strong>genes is <strong>of</strong> zoonotic pathogen.<br />

R E F E R E N C E<br />

A / P<br />

05<br />

R E F E R E N C E<br />

A / P<br />

06


R E F E R E N C E<br />

A / P<br />

07<br />

R E F E R E N C E<br />

A / P<br />

08<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Genotypic characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

from fresh leafy vegetables<br />

Warke, S. 1, Kalorey, D. R. 1, Umap, S. 1, Sonegaonkar, A. 1, Patil, V. 1, Kurkure, N V. 1<br />

and Barbuddhe, S. B. 2<br />

1. Nagpur Veterinary College, India<br />

2. ICAR Research Complex for Goa, India<br />

Listeria monocy<strong>to</strong>genes is a food borne pathogen responsible for listeriosis, an illness<br />

characterized by meningitis, encephalitis, abortion and septicemia in human<br />

beings. A <strong>to</strong>tal <strong>of</strong> 417 samples comprising eleven different types <strong>of</strong> fresh leafy vegetables<br />

were collected from supermarkets and local markets in and around Nagpur,<br />

Central India <strong>to</strong> detect L.monocy<strong>to</strong>genes. The vegetables included spinach (78),<br />

Fenugreek (82), Green Sorrel (22), Coriander (53), radish (57), amaranthus (30),<br />

sour greens (12), lettuce (33), Mint Leaves (43) and dill leaves (07). All samples<br />

were processed for isolation <strong>of</strong> L. monocy<strong>to</strong>genes by two step enrichment followed<br />

by plating on selective media. Confirmation <strong>of</strong> the isolates was on basis <strong>of</strong> biochemical<br />

characters, haemolysis on blood agar and Christie, Atkins, Munch Petersen<br />

test (CAMP). The isolates were also tested for the virulence associated<br />

genes namely, hlyA, plcA, actA, iap multiplex and prfA by PCR. A <strong>to</strong>tal <strong>of</strong> 31 (7.43 %)<br />

isolates <strong>of</strong> L. monocy<strong>to</strong>genes were recovered from spinach (11), fenugreek (9), coriander<br />

(3), radish (4), amaranthus (2) and mint leaves (2). All the genes were detected<br />

in seven strains, four genes (hlyA, actA iap and prfA) in three strains. The<br />

hlyA, actA and iap in six strains. Three genes (hlyA, iap and prfA) were detected in<br />

five strains. The hlyA and iap genes were detected in ten strains. Fourteen isolates<br />

were subjected <strong>to</strong> multiplex PCR serotyping and all belong <strong>to</strong> 4b, 4d and 4e<br />

serogroup. Contamination <strong>of</strong> vegetables with L.monocy<strong>to</strong>genes is a cause <strong>of</strong> concern<br />

as most <strong>of</strong> the vegetables are consumed raw as salad. Thus, t<strong>here</strong> is increased<br />

risk for transmission <strong>of</strong> listeriosis. T<strong>here</strong>fore proper washing and cleaning <strong>of</strong> fresh<br />

leafy vegetables is recommended at harvesting and consumption.<br />

Comprehensive appraisal <strong>of</strong> the exoproteome <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

by genomic and proteomic analyses<br />

Desvaux, M., Dumas, E., Chafsey, I., Chambon, C. and Hébraud, M.<br />

INRA, France<br />

Secreted proteins are the main <strong>to</strong>ols used by bacteria <strong>to</strong> interact with their environment<br />

and are relevant <strong>to</strong> the bacterial lifestyle. By definition, secreted proteins<br />

are proteins actively transported via secretion systems but they can have<br />

radically different final destinations. In monoderm (Gram-positive) bacteria, such<br />

secreted proteins can (i) anchor <strong>to</strong> the cy<strong>to</strong>plasmic membrane, (ii) associate with<br />

the cell wall, or (iii) be released in<strong>to</strong> the extracellular milieu and are the called exoproteins.<br />

The subset <strong>of</strong> proteins localized in the extracellular milieu constitutes<br />

the extracellular proteome, also called exoproteome. Listeria monocy<strong>to</strong>genes is a<br />

bacterial species both pathogenic and saprophytic, which exhibits a highly variable<br />

level <strong>of</strong> virulence from one strain <strong>to</strong> another and such biodiversity might express<br />

within the exoproteome. The present investigation aimed at developping <strong>of</strong><br />

a generic genomic strategy <strong>to</strong> predict exoproteins in monoderm bacteria as well as<br />

determining those differentially expressed within a panel <strong>of</strong> representative strains.<br />

Genomic analysis was performed following a novel rational bioinformatic approach<br />

<strong>to</strong> predict exoproteins, which expression was determined by proteomic<br />

analyses following 2-DE and identification by MALDI-TOF MS. The rational bioinformatic<br />

approach integrates the structural properties <strong>of</strong> the proteins potentially<br />

secreted as well as the respective secretion systems involved. Comparison <strong>of</strong> theoretical<br />

and experimental 2-DE maps provided a comprehensive picture <strong>of</strong> the exoproteome.<br />

Exoproteins were further discriminated between those found in all or<br />

only in a subset <strong>of</strong> L. monocy<strong>to</strong>genes strains. While the core exoproteome included<br />

proteins related <strong>to</strong> virulence and cell wall biogenesis, variation in the protein<br />

members <strong>of</strong> these categories constituted the variant exoproteome. The novel and<br />

rational in silico strategy <strong>here</strong> developed <strong>to</strong> predict exoproteins is adaptable and<br />

fully applicable <strong>to</strong> other monoderm bacteria. This proteomic analysis resulted in<br />

the first definition <strong>of</strong> the core and variant exoproteomes <strong>of</strong> this species and provides<br />

auxiliary insight in bacterial physiology.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Investigating differences in lineages <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

using comparative genomics<br />

McIlwham, S., Farber, J. and Pagot<strong>to</strong>, F.<br />

Bureau <strong>of</strong> Microbial Hazards, Canada<br />

Listeria monocy<strong>to</strong>genes causes foodborne listeriosis and is ubiqui<strong>to</strong>us in the environment.<br />

Strains are <strong>of</strong>ten classified in<strong>to</strong> lineages I (food); II (human illness); and<br />

III (food-production). It is not clearly unders<strong>to</strong>od how lineages develop a tropism<br />

for a particular niche. Using comparative genomics, we attempted <strong>to</strong> identify<br />

unique genetic attributes within lineages that contribute <strong>to</strong> the tropism <strong>of</strong> the<br />

pathogen. A mixed genome array, created from a genomic library <strong>of</strong> 15 L. monocy<strong>to</strong>genes<br />

isolates, was used <strong>to</strong> compare the genomes <strong>of</strong> different strains. An isogenic<br />

mutant for a glycoside hydrolase family 65 (GH65) was created based on microarray<br />

screening. Forty-five strains from different sources were studied in Mueller-<br />

Hin<strong>to</strong>n broth with ampicillin, vancomycin or acriflavine. Additionally, the growth<br />

rates <strong>of</strong> eleven strains in Listeria Enrichment Broth (LEB) were compared. A<br />

GH65 gene was found <strong>to</strong> be present in lineage II isolates but absent in lineage I isolates.<br />

A GH65 isogenic deletion mutant and other lineage I isolates showed a reduced<br />

ability <strong>to</strong> proliferate in the presence <strong>of</strong> vancomycin and ampicillin. When<br />

growth rates were compared in LEB, the GH65 deletion mutant behaved similarly<br />

<strong>to</strong> lineage I strains. The GH65 gene, present in lineage II strains, may provide a<br />

competitive advantage in the host environment. The difference in growth rate <strong>of</strong><br />

the lineage I and II strains in LEB may provide insight as <strong>to</strong> why certain serotypes<br />

are more readily isolated from food commodities.<br />

Au<strong>to</strong>lysis in Listeria monocy<strong>to</strong>genes – a proteomic approach<br />

Pin<strong>to</strong>, E. 1, Marques, N. 2, Andrew, P. W. 3 and Faleiro, M. L. 1<br />

1. Institu<strong>to</strong> de Biotecnologia e Bioengenharia, Centro de Biomedecina Molecular e Estrutural,<br />

Universidade do Algarve, Portugal<br />

2. BioFig, Universidade do Algarve, Portugal<br />

3. Department <strong>of</strong> Infection, Immunity and Inflammation, University <strong>of</strong> Leicester, UK<br />

Au<strong>to</strong>lysins play a role in many cellular processes including cell growth and division,<br />

cell wall turnover, motility and also pathogenicity. Several growth conditions<br />

may interfere with cell wall synthesis, leading <strong>to</strong> unrestricted activity <strong>of</strong> the<br />

au<strong>to</strong>lysins which can degraded the cell wall <strong>to</strong> levels that compromise the cell integrity.<br />

During our work differences between Listeria monocy<strong>to</strong>genes strains in<br />

the tendency <strong>to</strong> undergo au<strong>to</strong>lysis was found. The objective <strong>of</strong> the present study<br />

was <strong>to</strong> test the hypothesis that as cells enter stationary phase the pattern <strong>of</strong> protein<br />

expression in au<strong>to</strong>lytic and non-au<strong>to</strong>lytic strains is different. To achieve this,<br />

two L. monocy<strong>to</strong>genes strains (an au<strong>to</strong>lytic [C897] and a non-au<strong>to</strong>lytic [EGD])<br />

were grown in a defined medium at favourable (30 °C) and non-favourable (20 °C)<br />

au<strong>to</strong>lysis growth conditions. The bacterial cells and the supernatant samples were<br />

collected immediately before au<strong>to</strong>lysis. The samples were subjected <strong>to</strong> a proteome<br />

analysis. Remarkably at favourable au<strong>to</strong>lysis condition the au<strong>to</strong>lytic strain expressed<br />

a series <strong>of</strong> stress proteins and was depleted <strong>of</strong> 2-C-methyl-D-erythri<strong>to</strong>l 4phosphate<br />

cytidylyltransferase an essential component <strong>of</strong> the 2-C-methyl-D -<br />

erythri<strong>to</strong>l 4-phosphate (MEP) pathway. Injuries in this pathway may cause a<br />

negative impact on cell wall synthesis. In both proteomes <strong>of</strong> the non-au<strong>to</strong>lytic<br />

strain EGD and the au<strong>to</strong>lytic strain C897 in non favourable au<strong>to</strong>lyis conditions revealed<br />

a higher expression <strong>of</strong> proteins involved in the energy generation, amino<br />

acid and fatty acid metabolism. In the secre<strong>to</strong>me <strong>of</strong> C897 the p60 protein was<br />

over expressed at 30 °C but no significant differences were found between the<br />

strain C897 and EGD. However p60 levels were higher in the au<strong>to</strong>lytic strain intracellular<br />

proteome when grown at 20 °C. The levels <strong>of</strong> the au<strong>to</strong>lysin p45 were<br />

higher in the secre<strong>to</strong>me <strong>of</strong> C897 at 30 °C and significant differences between the<br />

two strains were found.<br />

R E F E R E N C E<br />

A / P<br />

09<br />

R E F E R E N C E<br />

A / P<br />

10


R E F E R E N C E<br />

A / P<br />

11<br />

R E F E R E N C E<br />

A / P<br />

12<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Role <strong>of</strong> flhA, cheR and motA in growth <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

at low temperature<br />

Mattila, M., Lindström, M., Somervuo, P. and Korkeala, H.<br />

University <strong>of</strong> Helsinki, Finland<br />

Temperature-dependent induction <strong>of</strong> flagella is a well-characterized phenomenon<br />

in L. monocy<strong>to</strong>genes. At 37 °C L. monocy<strong>to</strong>genes is virtually non-motile and the motility<br />

genes are down-regulated w<strong>here</strong>as at 30 °C and below the strains are highly flagellated<br />

and motile. However, the essentiality <strong>of</strong> increased flagellum production during<br />

growth at low temperatures is still unclear. To study this relationship, we<br />

prepared knock-out mutants <strong>of</strong> three motility and chemotaxis genes, flhA, cheR, and<br />

motA on L. monocy<strong>to</strong>genes strain EGD-e. We compared the expression levels <strong>of</strong> these<br />

genes in cultures grown at 3, 25 and 37 °C by using qRT-PCR. The role <strong>of</strong> these genes<br />

in the cold adaptation <strong>of</strong> L. monocy<strong>to</strong>genes was also studied with growth curve analysis,<br />

motility assays and electron microscopy. The relative expression <strong>of</strong> flhA, cheR<br />

and motA in EGD-e cells grown at 3 °C was significantly (p < 0.01) increased compared<br />

<strong>to</strong> 37 °C. At 3 °C the level <strong>of</strong> flhA and cheR transcripts was also significantly<br />

(p < 0.05) higher than at 25 °C. The growth curve analysis showed that in cultures<br />

grown at 3 °C both the growth rates and maximum optical densities at 600 nm were<br />

significantly (p < 0.001) lower for the mutant strains ∆flha, ∆cheR and ∆motA compared<br />

<strong>to</strong> the wild-type strain. At 37 and 25 °C we did not detect any significant differences<br />

between the wild-type strain and the mutants. The motility assays showed<br />

that ∆flhA and ∆motA strains were completely non-motile at all three temperatures<br />

w<strong>here</strong>as ∆cheR showed similar motility pattern <strong>to</strong> the wild-type at both 3 and 25 °C.<br />

At 37 °C the motility <strong>of</strong> ∆cheR mutant and the wild-type strain was negligible. The<br />

results suggest that flhA, cheR, and motA have a role in the cold <strong>to</strong>lerance <strong>of</strong> L. monocy<strong>to</strong>genes<br />

strain EGD-e, and that flagellum production and chemotaxis are linked <strong>to</strong><br />

cold adaptation <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

The effect <strong>of</strong> acetic acid (at pH 5.5) or benzoic acid (at neutral pH)<br />

on lipid composition and fluidity <strong>of</strong> Listeria monocy<strong>to</strong>genes membrane<br />

Ioannis, D. 1, Anita, B. 1, Eleni, S. 2 and Mastronicolis, S. 1<br />

1. Food Chemistry Labora<strong>to</strong>ry <strong>of</strong> Chemistry Dept, University <strong>of</strong> Athens, Panepistimiopolis Zografou,<br />

Athens, Greece<br />

2. National Hellenic Research Foundation, Institute <strong>of</strong> Organic and Pharmaceutical Chemistry, Athens,<br />

Greece<br />

Listeria monocy<strong>to</strong>genes is able <strong>to</strong> alter its membrane fatty acids (FA) composition or<br />

the composition <strong>of</strong> the polar group <strong>of</strong> the lipid molecules after a reduction in pH<br />

(due <strong>to</strong> Acid Tolerance Response, ATR). The lipid changes might play a protective<br />

role upon (acidic) stress decreasing the membrane fluidity and thus obstructing the<br />

entrance <strong>of</strong> the stressful compounds. The aim <strong>of</strong> our work was <strong>to</strong> investigate the<br />

modification <strong>of</strong> L. monocy<strong>to</strong>genes membrane lipids molecules (neutral, NL or polar,<br />

PL lipids) and also their FA composition in response <strong>to</strong> low (acetic acid) or neutral<br />

pH (benzoic acid) and their consequences on membrane fluidity. L. monocy<strong>to</strong>genes<br />

(DP-L1044, D. Portnoy, University <strong>of</strong> Pennsylvania) <strong>to</strong>tal lipids,TL extracted from<br />

stationary phase cells cultured (1L BHI broth, 30 °C) with low initial pH (5.5) by<br />

acetic acid, Lm AA (optical density OD 600 =0.210±0.031, 72h, n=12) or with neutral<br />

pH by benzoic acid (1g/L), Lm BA (OD 600 =0.682±0.013, 10h, n=9) and compared with<br />

those <strong>of</strong> optimal conditions cells, Lm control (OD 600 =0.816±0.032, 10h, n=8). The results<br />

revealed: i) Only acetic acid enhances the antimicrobial activity. ii) The sums<br />

<strong>of</strong> straight saturated chain FA(16:0 and 18:0) and unsaturated FA(16:1 and 18:1)<br />

were similarly increased, at the expense <strong>of</strong> branched chain FA(a-15:0 and a-17:0) <strong>of</strong><br />

each TL acidic adapted culture (increase <strong>of</strong> high-melting FA). iii) Singularly acetic<br />

acid adaptation in L. monocy<strong>to</strong>genes was correlated with a higher content (49.3 %<br />

among TL) <strong>of</strong> neutral lipids class than those <strong>of</strong> Lm control . Moreover, we correlated<br />

the lipid changes with thermodynamic analysis (Differential Scanning Calorimetry,<br />

DSC) <strong>of</strong> lipids in order <strong>to</strong> estimate experimentally the initial hypothesis, that both<br />

adaptation mechanisms (the high-melting point FA increase and the NL accumulation)<br />

would promote the decrease <strong>of</strong> membrane fluidity. The results revealed that<br />

in each culture Lm BA and Lm AA TL was increased the phase transition temperature,T<br />

c (5.3 °C and 3 °C) and enthalpy difference, ∆Η (10.2 % and 53.8 % respectively)<br />

than Lm control . The thermodynamic findings confirmed the above hypothesis<br />

<strong>of</strong> decreasing fluidity. The high ∆Η value <strong>of</strong> Lm AA (high energy amount <strong>to</strong><br />

change its lipids) is interpreted by NL accumulation.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Elucidation <strong>of</strong> the responses <strong>to</strong> weak acids in the human pathogen<br />

Listeria monocy<strong>to</strong>genes using gene microarrays<br />

O’Byrne, C. 1, Heavin, S. 1 and Morrissey, J. 2<br />

1. NUI Galway, Ireland<br />

2. University College Cork, Ireland<br />

Contamination <strong>of</strong> processed food with Listeria monocy<strong>to</strong>genes is a major human<br />

health risk. Food preservation regimes <strong>of</strong>ten include the use <strong>of</strong> weak organic acids<br />

<strong>to</strong> reduce microbial growth. Although the response <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> strong<br />

acids has been investigated extensively, t<strong>here</strong> little is known about how L. monocy<strong>to</strong>genes<br />

responds <strong>to</strong> weak acids, particularly at the molecular level. The objective<br />

<strong>of</strong> this study was <strong>to</strong> investigate the molecular response <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> five<br />

food-grade organic acids, acetic, lactic, sorbic, citric and benzoic acid. Careful<br />

growth experiments were carried out <strong>to</strong> determine the inhibi<strong>to</strong>ry concentrations<br />

<strong>of</strong> each <strong>of</strong> these acids on the growth <strong>of</strong> L. monocy<strong>to</strong>genes. Microarray and quantitative<br />

RT-PCR experiments were then performed <strong>to</strong> elucidate the molecular responses<br />

<strong>of</strong> the organism <strong>to</strong> weak acids. In <strong>to</strong>tal 222 genes were found <strong>to</strong> be differentially<br />

expressed in response <strong>to</strong> at least one <strong>of</strong> the weak acids. The temporal<br />

response <strong>of</strong> 9 <strong>of</strong> these genes <strong>to</strong> sudden exposure <strong>to</strong> lactic acid or citric acid was<br />

studied using RT-PCR. These data highlighted the distinct responses observed <strong>to</strong><br />

the 2 weak acids and identified genes that showed strong unique transcriptional<br />

responses (both up-regulation and down-regulation) <strong>to</strong> each. At a global level t<strong>here</strong><br />

was surprisingly little overlap between genes that responded <strong>to</strong> each acid. For example,<br />

the 7 genes differentially expressed in the presence <strong>of</strong> acetic acid were not<br />

influenced by any <strong>of</strong> the other acids. Together the data suggest that each organic<br />

acid influences the cell in a very specific way, a finding that suggests that each acid<br />

has a different inhibi<strong>to</strong>ry mode <strong>of</strong> action.<br />

The immunogenic surface protein IspC acts as an<br />

N-Acetylglucosaminidase in Listeria monocy<strong>to</strong>genes serotype 4b<br />

Ronholm, J.*<br />

University <strong>of</strong> Ottawa, Canada<br />

IspC is a surface protein isolated from Listeria monocy<strong>to</strong>genes serotype 4b which<br />

acts as an au<strong>to</strong>lysin and virulence fac<strong>to</strong>r. Several au<strong>to</strong>lysins including P60, Ami<br />

and Au<strong>to</strong> have been isolated from various L. monocy<strong>to</strong>genes serotypes but the role<br />

for their apparent functional overlap remains unknown. Further characterizing<br />

these au<strong>to</strong>lysins may help <strong>to</strong> gain insights in<strong>to</strong> their roles in Listeria biology. The<br />

novel au<strong>to</strong>lysin IspC remains understudied. A panel <strong>of</strong> 15 monoclonal antibodies<br />

(mAbs) directed against the surface <strong>of</strong> L. monocy<strong>to</strong>genes serotype 4b were generated,<br />

characterized as recognizing IspC and will facilitate further study <strong>of</strong> this protein.<br />

Based on serological assays with these mAbs we have determined that IspC<br />

expression is limited <strong>to</strong> certain L. monocy<strong>to</strong>genes serotypes including 4a, 4ab, 4b,<br />

and 4e, although, other serotypes have genes highly homologous <strong>to</strong> IspC. Sequence<br />

homology <strong>to</strong> other au<strong>to</strong>lysins predicted IspC functioned as an N-acetylmuramoyl-<br />

L-alanine amidase. However, mass spectrometry analysis <strong>of</strong> peptidoglycan (PG)<br />

hydrolysis with IspC has shown that IspC acts as an N-acetylglucosaminidase.<br />

Based on the site <strong>of</strong> PG cleavage, it was hypothesized that the amount <strong>of</strong> PG acetylation<br />

would affect IspC mediated PG hydrolysis. PgdA, the only one <strong>of</strong> three PG<br />

deacetylases predicted <strong>to</strong> localize <strong>to</strong> the cell surface, is required for de-acetylation<br />

<strong>of</strong> PG, which is assembled fully acetylated. To determine how acetylation affects<br />

IspC PG hydrolysis, a pgdA deletion mutant was created. The mutant PG<br />

(fully acetylated) was much more susceptible <strong>to</strong> hydrolysis by IspC than PG from<br />

the wild-type, as assessed by re-naturing SDS-PAGE. Currently we are making a<br />

pgdA complement <strong>to</strong> res<strong>to</strong>re PG deacetylation in the deletion mutant. Further investigation<br />

is being done on IspC hydrolysis <strong>of</strong> PG from the deletion mutant using<br />

mass spectrometry. This will provide insight in<strong>to</strong> the effects <strong>of</strong> the pgdA deletion<br />

on PG structure and IspC PG hydrolase activity.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

A / P<br />

13<br />

R E F E R E N C E<br />

A / P<br />

14


R E F E R E N C E<br />

A / P<br />

15<br />

R E F E R E N C E<br />

A / P<br />

16<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Listeria monocy<strong>to</strong>genes EGD chitinolytic activity is regulated by<br />

carbohydrates but also by the virulence regula<strong>to</strong>ry gene, PrfA<br />

Halberg Larsen, M., Leisner, J. J. and Ingmer, H.<br />

Faculty <strong>of</strong> Life Sciences, University <strong>of</strong> Copenhagen, Denmark<br />

Chitin, a highly insoluble polymer <strong>of</strong> N-acetyl-D-glucosamine (GlcNAc), is a major<br />

component <strong>of</strong> fungal cell walls and insect and crustacean exoskele<strong>to</strong>ns and is one <strong>of</strong><br />

the most abundant polymers in marine and terrestrial environments. Chitin hydrolysis<br />

by Listeria monocy<strong>to</strong>genes depends on two chitinase encoding genes chiA<br />

(lmo1883) and chiB (lmo105) and the aim <strong>of</strong> this study was <strong>to</strong> investigate their regulation.<br />

The production <strong>of</strong> the chitinases is substrate regulated and subjected <strong>to</strong><br />

catabolite control. Thus, chitin induces expression <strong>of</strong> both chitinases however chiA<br />

but not chiB is furthermore induced by the monomer GlcNAc. In growth medium<br />

supplemented with chitin and glucose the expression <strong>of</strong> both chitinases is repressed.<br />

Expression <strong>of</strong> chiA and chiB is growth phase dependent with higher expression in<br />

late exponential growth phase compared <strong>to</strong> early exponential phase. Chitinases expressed<br />

by bacterial pathogens have proven <strong>to</strong> be important not only for nutrient<br />

acquisition and environmental survival but also for infecting humans and animals.<br />

Interestingly, we found that the central L. monocy<strong>to</strong>genes virulence gene regula<strong>to</strong>r,<br />

PrfA, is required for the chitinolytic phenotype as chitinase activity was significantly<br />

reduced in ∆prfA mutant cells compared <strong>to</strong> the wild type cells. In agreement with<br />

this result northern blot analysis showed that the amounts <strong>of</strong> chiA and chiB transcripts<br />

were significantly lower upon induction by chitin in the ∆prfA mutant compared<br />

with the wild type. Furthermore, in contrast <strong>to</strong> the wild type, no chiA transcript<br />

could be detected in the mutant lacking PrfA during growth in medium<br />

supplemented with GlcNAc. Regulation <strong>of</strong> chitinolytic activity in L. monocy<strong>to</strong>genes<br />

is complex and the results obtained in this and other studies indicate that the biological<br />

role <strong>of</strong> this activity may not be limited <strong>to</strong> the external environment.<br />

Infectious dose curves for guinea pigs challenged with a Listeria<br />

monocy<strong>to</strong>genes epidemic clone strain and a strain carrying a naturallyoccurring<br />

virulence-attenuating mutation in inlA show a significant<br />

shift in median infectious dose<br />

Nightingale, K. 1, Van Stelten, A. 1, Simpson, J. M. 1, Chen, Y. 2, Scott, V. N. 3,<br />

Ross, W. H. 4, Whiting, R. C. 5 and Wiedmann, M. 6<br />

1. Colorado State University, USA<br />

2. Grocery Manufacturers Association, USA<br />

3. U.S. Food and Drug Administration, USA<br />

4. Health Canada<br />

5. Exponent, USA<br />

6. Cornell University, USA<br />

Listeria monocy<strong>to</strong>genes contains two subpopulations, including (i) epidemic clone<br />

(EC) strains, which have been linked <strong>to</strong> the majority <strong>of</strong> listeriosis outbreaks worldwide<br />

and are overrepresented among sporadic cases in some countries along with<br />

(ii) strains carrying mutations leading <strong>to</strong> a premature s<strong>to</strong>p codon (PMSC) in inlA,<br />

which are commonly isolated from ready-<strong>to</strong>-eat foods (approx. 45 %) but rarely associated<br />

with human disease (approx. 5 %). The virulence fac<strong>to</strong>r Internalin-A (InlA;<br />

encoded by inlA) binds certain is<strong>of</strong>orms <strong>of</strong> the cellular recep<strong>to</strong>r E-cadherin <strong>to</strong> facilitate<br />

crossing <strong>of</strong> the intestinal barrier by L. monocy<strong>to</strong>genes. The current<br />

FDA/FSIS/CDC L. monocy<strong>to</strong>genes risk assessment was developed with dose response<br />

data from murine challenge experiments, a model that fails <strong>to</strong> probe InlA<br />

mediated virulence due <strong>to</strong> the inability <strong>of</strong> InlA <strong>to</strong> bind the murine is<strong>of</strong>orm <strong>of</strong> E-cadherin.<br />

Guinea pigs, which express the human is<strong>of</strong>orm <strong>of</strong> E-cadherin that binds InlA,<br />

were intragastrically challenged with (i) a fully-invasive EC strain associated with a<br />

listeriosis outbreak or (ii) a strain carrying the most common inlA PMSC mutation.<br />

Dose-response curves for tissue infectivity were constructed with either a log-logistic,<br />

beta poisson, or exponential (for spleen data) fit <strong>to</strong> the raw individual and combined<br />

organ data. The log logistic and beta poisson models based on combined organ<br />

data showed an approx. 1.3 log10 CFU increase in the median infectious dose for the<br />

strain carrying a PMSC in inlA relative <strong>to</strong> the EC strain. Inclusion <strong>of</strong> strain effect<br />

significantly improved the ability <strong>of</strong> the model <strong>to</strong> explain the observed data, supporting<br />

a significant difference in tissue infectivity between EC and PMSC strains. L.<br />

monocy<strong>to</strong>genes strains thus show notable differences in infectious dose required <strong>to</strong><br />

establish an infection. Results from this work support the dose-response relationship<br />

for L. monocy<strong>to</strong>genes is strain-specific and will provide critical data for enhancement<br />

<strong>of</strong> existing and development <strong>of</strong> future risk assessments.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

MudPIT based proteomic analysis <strong>of</strong> alkaline adapted, environmentally<br />

persistent Listeria monocy<strong>to</strong>genes strains<br />

Nilsson, R. E., Ross, T. and Bowman, J. P.<br />

University <strong>of</strong> Tasmania, Australia<br />

Soluble protein expression pr<strong>of</strong>iles <strong>of</strong> environmentally persistent and sporadic<br />

Listeria monocy<strong>to</strong>genes strains in mid-exponential and stationary growth states<br />

at pH 7.3 and adapted <strong>to</strong> pH 8.5 were compared using multidimensional protein<br />

identification technology (MudPIT). Searches employed the X!Tandem algorithm<br />

and an in house data validation criteria was applied using the Trans Proteomic<br />

Pipeline (Version 3.4) and comparison against a decoy database. Qualitative and<br />

relative abundances were compared through functional on<strong>to</strong>logy (TIGR) and determination<br />

<strong>of</strong> spectral abundances. A <strong>to</strong>tal <strong>of</strong> 1661 proteins were identified following<br />

integration analysis. Significant differences in protein expression were observed<br />

for both growth state and culture condition. Increased presence and<br />

abundance <strong>of</strong> proteins associated with the cell wall, transport and binding, amino<br />

acid / c<strong>of</strong>ac<strong>to</strong>r biosynthesis and protein stabilization were identified. The study<br />

findings suggest alkaline adapted L. monocy<strong>to</strong>genes strains may use cy<strong>to</strong>plasmic<br />

acidification by surrogate pro<strong>to</strong>n sources (both endogenous and exogenous) as a<br />

means <strong>of</strong> persisting in high pH environments. Furthermore, proteins associated<br />

with stabilization <strong>of</strong> cellular processes and cell wall modification may be aiding<br />

in minimizing cellular damage under these conditions. This trend was identified in<br />

both growth states; however it was more pronounced in exponential phase. No<br />

significant difference was observed between the L. monocy<strong>to</strong>genes strains studied.<br />

The results <strong>of</strong> this work support the notion that exposure <strong>to</strong> conditions favouring<br />

alkaline homeostasis may be an important contribu<strong>to</strong>r <strong>to</strong> the development <strong>of</strong> persistent<br />

L. monocy<strong>to</strong>genes, independent <strong>of</strong> strain.<br />

RpoN, the alternative sigma fac<strong>to</strong>r, is associated with the growth phase<br />

transition and pathogenesis in Listeria monocy<strong>to</strong>genes<br />

Okada, Y., Suzuki, H., Monden, S., Igimi, S. and Okada, N.<br />

Ministry <strong>of</strong> Health, Labour and Welfare, Japan<br />

Listeria monocy<strong>to</strong>genes has 5 types <strong>of</strong> sigma fac<strong>to</strong>r. RpoN, an alternative sigma fac<strong>to</strong>r,<br />

is involved in the bacteriocin-resistance and the osmo<strong>to</strong>lerance in this bacterium,<br />

however, its’ function remains almost unknown. We found that the rpoN<br />

deletion mutant showed higher growth than its parental strain at the stationary<br />

phase. To identify the RpoN regulated genes which are associated with the growth<br />

phase transition, DNA microarray analysis was performed. From L. monocy<strong>to</strong>genes<br />

strain EGD and its rpoN in-frame deletion mutant, <strong>to</strong>tal RNA samples were isolated<br />

at early exponential and stationary phase. The gene expression pr<strong>of</strong>iles in<br />

both strains at each growth phase were compared, and in the results, especially, 27<br />

genes associated with oxidoreductase showed the different patterns <strong>of</strong> expression<br />

between two strains. So we examined the survival <strong>of</strong> these strains after exposure <strong>to</strong><br />

the sub-lethal concentration <strong>of</strong> hydrogen peroxide. Both strains showed the similar<br />

levels <strong>of</strong> <strong>to</strong>lerance against 100 mM hydrogen peroxide at stationary phase. However,<br />

the mutant was significantly <strong>to</strong>lerant <strong>to</strong> hydrogen peroxide than EGD at the<br />

mid exponential phase. From the result <strong>of</strong> electrophoresis, the chromosomal DNA<br />

was degraded after exposure <strong>to</strong> hydrogen peroxide only in EGD, but not in the mutant.<br />

Next, we examined the pathogenicity <strong>of</strong> the rpoN mutant. The mutant<br />

showed the higher growth in J774 cells and the peri<strong>to</strong>neal ad<strong>here</strong>nt cells from<br />

BALB/c mice. In spleen and liver <strong>of</strong> mouse, the mutant was able <strong>to</strong> grow at the<br />

similar level <strong>of</strong> EGD, however, the growth was slowly. These results suggest that the<br />

rpoN mutant is able <strong>to</strong> grow higher than the parental strain at the stationary phase<br />

because <strong>of</strong> the increased DNA stability against the active oxygen accumulated in<br />

the bacterial cells as a result <strong>of</strong> the respiration during the exponential growth, and<br />

RpoN is associated with the pathogenicity <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

R E F E R E N C E<br />

A / P<br />

17<br />

R E F E R E N C E<br />

A / P<br />

18


R E F E R E N C E<br />

A / P<br />

19<br />

R E F E R E N C E<br />

A / P<br />

20<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Cellular lipid fatty acid pattern differences between reference and<br />

ice-cream isolate <strong>of</strong> Listeria monocy<strong>to</strong>genes as response <strong>to</strong> cold stress<br />

Anita, B., Ioannis, D. and Mastronicolis, S.<br />

Food Chemistry Labora<strong>to</strong>ry, Department <strong>of</strong> Chemistry, National University <strong>of</strong> Athens, Greece<br />

One <strong>of</strong> the salient features <strong>of</strong> L. monocy<strong>to</strong>genes as a food-borne pathogen is its<br />

ability <strong>to</strong> grow at low temperatures, allowing refrigeration at 5 °C <strong>to</strong> act as an effective<br />

enrichment for the organism. T<strong>here</strong> is little information about its ability <strong>to</strong><br />

withstand in frozen foods. L.monocy<strong>to</strong>genes is characterized by > 85 % branchedchain<br />

fatty acids (FA), anteiso-15:0 (a-15:0) and anteiso-17:0 (a-17:0). Cells grown in<br />

the cold (5 °C) contain significantly less a-17:0 than those grown at higher temperatures.<br />

The purpose was <strong>to</strong> investigate the effect <strong>of</strong> ice-cream s<strong>to</strong>rage in<br />

L.monocy<strong>to</strong>genes FA pr<strong>of</strong>ile, after inoculation <strong>of</strong> pilot ice-cream with reference<br />

strain DP-L1040 (D. Portnoy, University <strong>of</strong> Pennsylvania). The inoculation was<br />

performed with reference strain Lm ref , (BHI, 30 °C, 8h, OD 600 =0.8) in order <strong>to</strong><br />

yield 6.30 log cfu/g ice-cream. Cells <strong>of</strong> L. monocy<strong>to</strong>genes isolated from inoculated<br />

ice-cream, (Lm ice ), after 6 month <strong>of</strong> s<strong>to</strong>rage and cells <strong>of</strong> Lm ref were grown in BHI<br />

broth, until late exponential phase at two different temperatures, at 30 °C (10h,<br />

OD 600 =0.8) and at 5°C (8d, OD 600 =0.6). Each culture (Lm ice30°C , Lm ice5°C ,<br />

Lm ref30°C and Lm ref5°C ) <strong>to</strong>tal lipids were extracted and FA methyl esters were analyzed.<br />

Our results show that significant differences exist between the FA pr<strong>of</strong>ile <strong>of</strong><br />

Lm ref30°C and Lm ice30°C . The Lm ice30°C was characterized by a-15:0 and a-17:0 FA.<br />

Additionally, increase <strong>of</strong> straight chain 18:0 FA, (from 0.7 <strong>to</strong> 3.9 %) was observed in<br />

cells <strong>of</strong> ice-cream isolate. T<strong>here</strong> was also an increase <strong>of</strong> unsaturated FA. The<br />

Lm ice5°C showed an a-15:0/a-17:0 ratio <strong>of</strong> 11.2, while Lm ref5°C was characterized<br />

by significantly higher ratio <strong>of</strong> 19.3. Furthermore, t<strong>here</strong> was a decrease (3.2 fold) <strong>of</strong><br />

∑branched/∑straight chain FA ratio in ice-cream isolate compared <strong>to</strong> reference<br />

strain. These results showed that L. monocy<strong>to</strong>genes strain origin needs <strong>to</strong> be considered<br />

when physiological studies are performed with this bacterium. The food<br />

matrix and environmental conditions can influence the FA pattern and L. monocy<strong>to</strong>genes<br />

survival. A greater understanding <strong>of</strong> cold adaptation mechanism may<br />

<strong>of</strong>fer methods for controlling L. monocy<strong>to</strong>genes in chilled and frozen foods.<br />

Role <strong>of</strong> the dihydroxyace<strong>to</strong>ne metabolism in the resistance <strong>of</strong> Listeria<br />

innocua <strong>to</strong> pediocin<br />

Milohanic, E.<br />

INRA, France<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive bacterium, the causative agent <strong>of</strong> listeriosis,<br />

which affects humans and animals. This bacterium which is a recurring contaminant<br />

<strong>of</strong> the food chain represents a major problem for the food industry. We<br />

were interested in the dha system <strong>of</strong> Listeria innocua that involves the general PTS<br />

enzymes, the main transport and phosphorylation system <strong>of</strong> carbohydrates, which<br />

also occurs in L. monocy<strong>to</strong>genes. It catalyses the intracellular PEP-dependent phosphorylation<br />

<strong>of</strong> dihydroxyace<strong>to</strong>ne (Dha). In silico studies allowed us <strong>to</strong> hypothesize<br />

that the last gene <strong>of</strong> the dha system, pedB, could encode resistance <strong>to</strong> pediocin, a<br />

bacteriocin produced by the lactic acid bacterium Pediococcus acidilactici. We<br />

sought <strong>to</strong> understand why this gene is associated with the dha system. We first confirmed<br />

the operon organization <strong>of</strong> this system. Quantitative PCR experiments revealed<br />

that the expression <strong>of</strong> the dha system seems <strong>to</strong> be partially constitutive. Elevated<br />

dha expression was observed in a dhaR mutant, which lacks the repressor <strong>of</strong><br />

this operon. We also showed that the wild-type strain <strong>of</strong> L. innocua is sensitive <strong>to</strong><br />

pediocin while the dhaR mutant is resistant. The pedB mutant is under investigation.<br />

In addition, we purified the five proteins involved in Dha phosphorylation<br />

and could reconstitute in vitro these metabolic steps. Interestingly the dha operon<br />

encodes many enzymes involved in the pen<strong>to</strong>se phosphate pathway especially the<br />

Tpi-like protein. These results would suggest a link between Dha metabolism, pen<strong>to</strong>se<br />

phosphate pathway and virulence in L. monocy<strong>to</strong>genes.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Glucose transport system in Listeria monocy<strong>to</strong>genes and their impact<br />

on virulence gene expression<br />

Moussan Ake, F.<br />

INRA, France<br />

In Listeria monocy<strong>to</strong>genes at least two phosphoenolpyruvate (PEP): carbohydrate<br />

phosphotransferase systems (PTS) <strong>of</strong> the mannose class and also two non-PTS<br />

permeases can transport glucose. Mutants lacking EIIAB (Lmo0096) or EIIC<br />

(Lmo0097) <strong>of</strong> the glucose/mannose PTS grew slightly slower on minimal medium<br />

containing up <strong>to</strong> 10 mM glucose and consumed glucose 3-times slower than the<br />

wild-type strain. A lmo0784 mutant (EIIA, mannose type PTS) exhibited an extended<br />

lag phase before it started growing on glucose and glucose consumption<br />

was slowed 3.5-fold. All PTS transporters are inactive in a ptsI mutant, which nevertheless<br />

grew at high glucose concentrations (> 15 mM) after an extended lag<br />

phase. Two GlcU-like non-PTS glucose transporters (Lmo0169 and Lmo0176)<br />

probably catalyze glucose uptake in the DptsI mutant. Expression <strong>of</strong> either one <strong>of</strong><br />

them in an Escherichia coli DptsHIcrr mutant res<strong>to</strong>red glucose utilization. The efficient<br />

utilization <strong>of</strong> glucose affects the PrfA-dependent expression <strong>of</strong> L. monocy<strong>to</strong>genes<br />

virulence genes, including hly. Nevertheless, inactivation <strong>of</strong> the glcU genes<br />

in a strain carrying a Phly-gus fusion had no effect on PrfA activity. However, deletion<br />

<strong>of</strong> the gene for EIIAB Glc/Man caused 3- <strong>to</strong> 5-fold elevated gus expression,<br />

w<strong>here</strong>as deletion <strong>of</strong> EIIC Glc/Man had only a slight effect. b-Glucuronidase activity<br />

in the EIIAB Glc/Man mutant was also elevated when it was grown on solid minimal<br />

medium containing 5 or 10 mM glucose. A similar behaviour was observed for<br />

the lmo0784 mutant, which in addition exhibits very low man operon expression.<br />

The highest gus expression was observed in the DptsI mutant grown at glucose<br />

concentrations between 25 and 50 mM.<br />

Antimicrobial susceptibilities <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

in Japan<br />

Monden, S., Okutani, A., Suzuki, H., Asakura, H., Nakama, A., Igimi, S., Okada, Y.<br />

and Maruyama, T.<br />

Japan<br />

Antibiotics, such as ampicillin and co-trimoxazole, are usually used as the first<br />

choice for treatment <strong>of</strong> listeriosis. However, the antimicrobial resistant strains <strong>of</strong><br />

Listeria monocy<strong>to</strong>genes are <strong>of</strong>ten isolated in many countries. In this study, in vitro<br />

antimicrobial susceptibility <strong>of</strong> 232 L. monocy<strong>to</strong>genes strains (101 isolates from patients<br />

<strong>of</strong> listeriosis and 131 isolates from food and food-related environment) obtained<br />

in Japan from 1974 <strong>to</strong> 2008 were examined their susceptibilities against 6<br />

kinds <strong>of</strong> antibiotics by plate dilution method. According <strong>to</strong> the breakpoint from the<br />

guideline <strong>of</strong> National Comittee for Chemical Labora<strong>to</strong>ry Standard, all strains tested<br />

<strong>here</strong> were susceptible <strong>to</strong> gentamicin and kanamycin, and all clinical isolates<br />

showed susceptibility <strong>to</strong> ampicillin and erythromycin. One strain from beef was resistant<br />

<strong>to</strong> ampicillin, and the another one from pork showed resistance <strong>to</strong> 3 kinds <strong>of</strong><br />

drug, chloramphenicol, erythromycin and enr<strong>of</strong>loxacin. 56.4 % <strong>of</strong> clinical isolates<br />

and 37.8 % <strong>of</strong> food and environmental isolates were resistant <strong>to</strong> enr<strong>of</strong>loxacin, and<br />

the rest <strong>of</strong> isolates showed intermediate phenotype <strong>to</strong> this antibiotics. The rates <strong>of</strong><br />

enr<strong>of</strong>loxacin-resistance were very high in isolates belonging <strong>to</strong> serotype 1/2a<br />

(63.3 %) and 1/2c (75 %) compared with those <strong>of</strong> 1/2b (41.8 %) and 4b (38 %). In this<br />

study, most <strong>of</strong> clinical isolates were susceptible <strong>to</strong> antibiotics except enr<strong>of</strong>loxacin.<br />

On the other hand, some strains from food and environment showed resistance or<br />

intermediate <strong>to</strong> ampicillin, chloramphenicol or erythromycin. Furthermore, all<br />

strains were more than intermediate <strong>to</strong> enr<strong>of</strong>loxacin, a kind <strong>of</strong> quinolone that used<br />

for treatment <strong>of</strong> animal infection in Japan. These results suggest that the origin <strong>of</strong><br />

listerial isolates in involved in their antibiotic susceptibility.<br />

R E F E R E N C E<br />

A / P<br />

21<br />

R E F E R E N C E<br />

A / P<br />

22


R E F E R E N C E<br />

A / P<br />

23<br />

R E F E R E N C E<br />

A / P<br />

24<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Influence <strong>of</strong> sub-lethal concentrations <strong>of</strong> disinfectants<br />

on Listeria monocy<strong>to</strong>genes adhesion and invasion in Caco-2 cells<br />

Gaedt Kastbjerg, V. 1, Halberg Larsen, M. 2, Ingmer, H. 2 and Gram, L. 1<br />

1. National Institute <strong>of</strong> Food Research, Technical University <strong>of</strong> Denmark<br />

2. Faculty <strong>of</strong> Life Sciences, University <strong>of</strong> Copenhagen, Denmark<br />

Listeria monocy<strong>to</strong>genes is frequently detected in the food processing environment,<br />

w<strong>here</strong> it, ideally, must be exposed <strong>to</strong> disinfectants daily. However, L. monocy<strong>to</strong>genes<br />

is not always eliminated by the disinfection process. One reason could be<br />

that the bacteria are only exposed <strong>to</strong> sub-lethal concentrations <strong>of</strong> the disinfectant.<br />

We have recently shown that sub-lethal concentrations <strong>of</strong> disinfectants used in<br />

the food industry affect virulence gene expression on transcript level in L. monocy<strong>to</strong>genes,<br />

and the effect depend on the active components <strong>of</strong> the disinfectants.<br />

The aim <strong>of</strong> the present study was <strong>to</strong> determine if disinfectants used routinely in<br />

the food industry affect the virulence <strong>of</strong> this pathogen studied in a cell-model.<br />

Two different disinfectants were used. Compound 1 contains peracetic acid and<br />

hydrogen peroxide as the active ingredients and reduces the expression <strong>of</strong> virulence<br />

genes in L. monocy<strong>to</strong>genes w<strong>here</strong>as Compound 2 contains quaternary ammonium<br />

compounds (QAC) and induces virulence gene expression. L. monocy<strong>to</strong>genes<br />

EGD was exposed <strong>to</strong> the two disinfectants for one hour in a non-inhibiting<br />

and a sub-lethal concentration, and subsequently the bacterial cells were added<br />

<strong>to</strong> a monolayer <strong>of</strong> Caco-2 cells. Bacterial adhesion and invasion were moni<strong>to</strong>red.<br />

Exposure <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> the sub-lethal concentration (0.0031 %) <strong>of</strong> the<br />

QAC disinfectant significantly increased adhesion <strong>to</strong> Caco-2 cells as compared <strong>to</strong><br />

bacteria exposed <strong>to</strong> water (control) and resulted in a slightly higher invasion. No<br />

effect was observed <strong>of</strong> the non-inhibiting concentration <strong>of</strong> the QAC disinfectant<br />

(0.0016 %). In contrast, the adhesion was unaffected and the invasion slightly decreased<br />

after exposure <strong>to</strong> the non-inhibiting concentration <strong>of</strong> 0.125 % <strong>of</strong> the peracetic<br />

disinfectant as compared <strong>to</strong> bacteria exposed <strong>to</strong> water. On-going studies<br />

will evaluate how long term exposure <strong>to</strong> disinfectants will affect adhesion and invasion<br />

<strong>to</strong> Caco-2 cells and these data will also be discussed.<br />

The SOS response in Listeria monocy<strong>to</strong>genes – a stress<br />

survival mechanism<br />

Kiil Nielsen, P., Zahle Andersen, A. and Haahr Kallipolitis, B.<br />

University <strong>of</strong> Southern Denmark<br />

The SOS response is a functionally conserved DNA repair system found in a variety<br />

<strong>of</strong> bacteria. The SOS response involves two central regula<strong>to</strong>ry components, RecA<br />

and LexA, which coordinate the expression <strong>of</strong> target genes leading <strong>to</strong> arrest <strong>of</strong> cell<br />

division, DNA repair and mutagenesis. His<strong>to</strong>rically, the SOS response is linked <strong>to</strong><br />

the presence <strong>of</strong> single stranded DNA and stalled replication forks induced by UV radiation.<br />

However several other stress conditions, which are not expected <strong>to</strong> cause<br />

massive DNA damage, have been shown <strong>to</strong> induce SOS response e.g. mild heat<br />

treatment, antibiotics, ethanol and salt stress. The SOS response <strong>of</strong> the food borne<br />

pathogen Listeria monocy<strong>to</strong>genes is presently not well unders<strong>to</strong>od. We find it likely<br />

that the SOS response is important for the growth and survival <strong>of</strong> this pathogen in<br />

contaminated foods and during infection. It is t<strong>here</strong>fore <strong>of</strong> great interest <strong>to</strong> investigate<br />

this response in L. monocy<strong>to</strong>genes in<strong>to</strong> more details. To this end, we focused<br />

our studies on the au<strong>to</strong> regulated repressor protein LexA, the genes targeted by this<br />

regula<strong>to</strong>r, and the environmental signals leading <strong>to</strong> activation <strong>of</strong> the SOS response<br />

in L. monocy<strong>to</strong>genes. LexA regulated gene expression is analyzed by comparing expression<br />

pr<strong>of</strong>iles <strong>of</strong> WT L. monocy<strong>to</strong>genes with ∆lexA and lexA* strains expressing a<br />

non functional and a constitutive active LexA, respectively. In order <strong>to</strong> further elucidate<br />

the regula<strong>to</strong>ry mechanisms and timing <strong>of</strong> the SOS response, we construct<br />

promoter-reporter gene fusions, which report the size and duration <strong>of</strong> activation <strong>of</strong><br />

LexA controlled protein expression. As reporter genes we employ both lacZ and<br />

degradation tagged gfp. We thus expect <strong>to</strong> be able <strong>to</strong> moni<strong>to</strong>r time resolved changes<br />

in SOS gene expression and thus come closer <strong>to</strong> the apparently versatile role <strong>of</strong> this<br />

stress response mechanism in Listeria monocy<strong>to</strong>genes.


Acid shock triggers heavy metal de<strong>to</strong>xification<br />

in Listeria monocy<strong>to</strong>genes<br />

Müller, S., Neuhaus, K. and Sc<strong>here</strong>r, S.<br />

Technische Universität München, Germany<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

L. monocy<strong>to</strong>genes is a food borne pathogen ubiqui<strong>to</strong>us in the environment. The<br />

bacterium is exposed <strong>to</strong> low pH in various ecological niches, such as the s<strong>to</strong>mach,<br />

macrophages or acidic soil. Acidic conditions are also found in ecological niches associated<br />

with food production. In this work, the acid shock response at pH 5 for<br />

15 min, 30 min, 45 min, 60 min and 120 min and the acid adaptation at pH 5.2<br />

(growth <strong>to</strong> mid log-phase) was studied at the transcriptional level using microarrays.<br />

Besides transcriptional induction <strong>of</strong> well characterized genes which are involved<br />

in acid <strong>to</strong>lerance, an up regulation was also found for genes whose products<br />

are potentially participating in heavy metal de<strong>to</strong>xification, such as ion export<br />

proteins (heavy metal exporting ATPase: lmo0641, lmo1852-1854; cation efflux<br />

protein: lmo2231, lmo2231) or heavy metal reduction proteins (arsenate reductase:<br />

lmo2230). Furthermore, it was demonstrated that, after induction <strong>of</strong> the acid<br />

<strong>to</strong>lerance response (60 min, pH 6) L. monocy<strong>to</strong>genes exhibits a higher resistance<br />

against various heavy metal ions, e.g. against Cd II, Hg II, and Ni II; and especially<br />

against Cu II and As V. Acidic pH appears <strong>to</strong> serve as a trigger not only for the induction<br />

<strong>of</strong> genes involved in acid <strong>to</strong>lerance itself, but also for the induction <strong>of</strong><br />

genes involved in heavy metal resistance. Since metals become more bioaccessible<br />

at low pH, especially in soil, acid solubilized heavy metals might reach noxious<br />

doses in certain environments and, t<strong>here</strong>fore, de<strong>to</strong>xification genes may be up regulated.<br />

Ongoing analysis using deletion mutants will help <strong>to</strong> elucidate the role <strong>of</strong><br />

de<strong>to</strong>xification under acidic growth conditions.<br />

Listeria monocy<strong>to</strong>genes mutants defective in growth at 5 °C<br />

and in high salt environment<br />

Burall, L., Laksanalamai, P. and Datta, A.<br />

US Food and Drug Administration, USA<br />

Listeria monocy<strong>to</strong>genes can survive and grow in refrigerated temperature and in<br />

the presence <strong>of</strong> high salt. In an effort <strong>to</strong> better understand the associated mechanisms,<br />

a library <strong>of</strong> 5280 transposon mutants <strong>of</strong> LS411, an isolate from the Jalisco<br />

cheese outbreak, were screened for their ability <strong>to</strong> grow in brain heart infusion<br />

(BHI) broth at 5 °C or in the presence <strong>of</strong> 7 % NaCl. Two mutants with altered<br />

growth pr<strong>of</strong>iles have been identified. 63-C8 showed a significant reduction in<br />

growth in BHI at 5 °C. Additionally, the mutant had reduced growth in the presence<br />

<strong>of</strong> 9 % but not 7 % NaCl. Sequencing located the transposon between two divergently<br />

transcribed genes, iap and one encoding a putative SecA subunit. qPCR<br />

revealed a substantial reduction in the expression <strong>of</strong> iap in 63-C8 relative <strong>to</strong> LS411,<br />

though a slight increase in expression was observed for the putative SecA subunit<br />

gene. 36-F11 showed no growth reduction at 5 °C but had attenuated growth in the<br />

presence <strong>of</strong> 7 % NaCl. The salt attenuation was exacerbated at 9 % NaCl, which resulted<br />

in reduced colony counts despite apparent increases in biomass. Preliminary<br />

microscopy work showed that the majority <strong>of</strong> cells in these cultures are “live”,<br />

indicating the possible presence <strong>of</strong> either a viable but non-culturable state or increased<br />

avidity in the bacterial associations within the clumps. Sequencing <strong>of</strong> the<br />

insertion region <strong>of</strong> 36-F11 identified the transposon between two divergently transcribed<br />

genes. The insertion site is 36bp upstream <strong>of</strong> a gene encoding a glycine<br />

betaine/L-proline ABC transporter, which is down-regulated in the mutant while<br />

the divergently transcribed putative mechanosensitive ion channel appeared<br />

slightly upregulated. Further studies are being conducted <strong>to</strong> identify the mechanisms<br />

underlying these phenotypes.<br />

R E F E R E N C E<br />

A / P<br />

25<br />

R E F E R E N C E<br />

A / P<br />

26


R E F E R E N C E<br />

A / P<br />

27<br />

R E F E R E N C E<br />

A / P<br />

28<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Revival <strong>of</strong> 5,000 Listeria strains from Seeliger’s his<strong>to</strong>rical collection<br />

with a semi-au<strong>to</strong>mated microbiological pipeline<br />

Haase, J. 1, H<strong>of</strong>, H. 2 and Achtman, M. 1<br />

1. Science Foundation Ireland<br />

2. University <strong>of</strong> Heidelberg, Mannheim, Germany<br />

Pr<strong>of</strong>. Seeliger from the University <strong>of</strong> Würzburg, Germany, collected more than<br />

7,000 Listeria strains, globally, between the 1920s and the 1980s from humans, animals,<br />

food and the environment. Within this project, this collection, the “Special<br />

Listeria Culture Collection” (SLCC), has been recultured and frozen at -80 °C.<br />

Strain information has been transcribed from the original notebooks and is now<br />

s<strong>to</strong>red in an SQL-based database (Bionumerics, Applied Maths) and the physical<br />

strain locations are s<strong>to</strong>red in a LIM system (ItemTracker). A large collection such<br />

as this required the set up <strong>of</strong> a comprehensive “strain pipeline”. This ensures correct<br />

cataloguing, efficient s<strong>to</strong>rage and retrieval <strong>of</strong> cultures. This was achieved via<br />

cus<strong>to</strong>m scripts (Python) that allow communication and archiving <strong>of</strong> data obtained<br />

from the initial culturing <strong>to</strong> isolation <strong>of</strong> DNA and downstream phylogenetic analyses.<br />

The collection now provides an exceptional opportunity <strong>to</strong> investigate the<br />

population structure and the evolutionary his<strong>to</strong>ry <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

Two point mutations are responsible for the lack <strong>of</strong> glycosidic substition<br />

in cell wall teichoic acids in Listeria monocy<strong>to</strong>genes serovar “7”<br />

Eugster, M. R. 1, Huwiler, S. 2, Morax, L. 1 and Loessner, M. J. 1<br />

1. Institute <strong>of</strong> Food, Nutrition and Health, ETH Zurich, Switzerland<br />

2. Institute <strong>of</strong> Microbiology, ETH Zurich, Switzerland<br />

The currently defined six Listeria species are further differentiated in<strong>to</strong> at least 15<br />

distinct serovars based on serological reactions with specific antisera directed<br />

against O and, <strong>to</strong> a lesser extent, H antigens. Some L. monocy<strong>to</strong>genes strains (such<br />

as WSLC 1034) have previously been assigned <strong>to</strong> serovar “7”. These cells contain<br />

ribi<strong>to</strong>l phosphate wall teichoic acids (WTA) similar <strong>to</strong> serovars 1/2 and 3, but without<br />

glycosyl substition. Purified WTA polymers and associated sugars were analyzed<br />

by a novel technique using electrospray ionization tandem mass spectrometry<br />

(ESI-MS/MS). In addition, WTA carbohydrates were identified using<br />

fluorescently labeled N-acetylglucosamine (GlcNAc) binding proteins (CBD-P35)<br />

and Listeria phage A118 which requires rhamnose-substituted WTAs for adsorption.<br />

We found that the absence <strong>of</strong> rhamnose and GlcNAc in WTAs <strong>of</strong> strain 1034 is<br />

based on two single point mutations only. The loss <strong>of</strong> GlcNac is due <strong>to</strong> a mutation<br />

(P300L) in the protein <strong>of</strong> gene lmo2550, which is involved in wall teichoic acid<br />

decoration with GlcNAc. The second mutation is a single base deletion and produces<br />

a frameshift at position 165 in gene lmo1083, whose product is involved in<br />

rhamnose biosynthesis, resulting in a truncated enzyme and lack <strong>of</strong> rhamnose in<br />

WTA. Deletion mutants constructed in a serovar 1/2a background (EGDe<br />

Dlmo2550 Dlmo1083) resulted in loss <strong>of</strong> the rhamnose and GlcNAc WTA components.<br />

Complementation with wt lmo2550 and lmo1083 in the EGDe mutant, in<br />

serovar “7” strain 1034, and in other mutants featuring a lack <strong>of</strong> sugar substitution<br />

not only res<strong>to</strong>red the rhamnose and GlcNAc content in EGDe, but also produced<br />

a serovar 1/2 phenotype in the “serovar 7” strain and other mutants lacking<br />

GlcNAs from their WTA. These findings suggest that Listeria serovar “7” strains (at<br />

least WSLC 1034) are likely mutants originating from a serotype 1/2 background,<br />

and that Listeria monocy<strong>to</strong>genes serovar “7” may not exist.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

High-throughput genome sequencing <strong>of</strong> two Listeria monocy<strong>to</strong>genes<br />

clinical isolates during a large foodborne outbreak<br />

Gilmour, M. 1, Graham, M. 1, Van Domselaar, G. 1, Tyler, S. 1, Kent, H. 1, Trout-Yakel, K. 1,<br />

Larios, O. 1, Allen, V. 2, Lee, B. 3, Nadon, C. 1 and Kearney, A. 1<br />

1. Public Health Agency <strong>of</strong> Canada<br />

2. Ontario Agency for Health Protection and Promotion, Canada<br />

3. Canadian Food Inspection Agency, Canada<br />

A large, multi-province outbreak <strong>of</strong> listeriosis associated with ready-<strong>to</strong>-eat meat<br />

products contaminated with Listeria monocy<strong>to</strong>genes serotype 1/2a occurred in<br />

Canada in 2008. Subtyping <strong>of</strong> outbreak-associated isolates using pulsed-field gel<br />

electrophoresis (PFGE) revealed two similar but distinct AscI PFGE patterns.<br />

Whole genome sequencing <strong>of</strong> two L. monocy<strong>to</strong>genes isolates was completed using<br />

the Roche GS FLX platform <strong>to</strong> rapidly determine the genome sequence <strong>of</strong> the<br />

primary outbreak strain and <strong>to</strong> investigate the genetic diversity associated with a<br />

change <strong>of</strong> a single restriction enzyme fragment observed by PFGE. The chromosomes<br />

were collinear, but differences included 28 single nucleotide polymorphisms<br />

(SNPs) and three indels, including a 33 kbp prophage that accounted for<br />

the difference in PFGE pattern. The distribution <strong>of</strong> these traits was assessed<br />

within clinical, environmental and food isolates associated with the outbreak, and<br />

this comparison indicated that three distinct, but highly related strains may have<br />

been involved in this nationwide outbreak. Notably, these two isolates were found<br />

<strong>to</strong> harbor a 50 kbp putative mobile genomic island encoding translocation and efflux<br />

functions that has not been observed in other Listeria genomes. Highthroughput<br />

genome sequencing t<strong>here</strong>fore provided a more detailed real-time assessment<br />

<strong>of</strong> genetic traits characteristic <strong>of</strong> the outbreak strains than could be<br />

achieved with routine subtyping methods. These data allowed us <strong>to</strong> determine<br />

evolutionary lineages and unequivocally define the full breadth <strong>of</strong> genetic variation<br />

between two subtype variants identified by the internationally standardized<br />

PulseNet PFGE typing method. This study confirms that the latest generation <strong>of</strong><br />

DNA sequencing technologies can be applied during high priority public health<br />

events, and labora<strong>to</strong>ries need <strong>to</strong> prepare for this inevitability and assess how <strong>to</strong><br />

properly analyze and interpret whole genome sequences in the context <strong>of</strong> molecular<br />

epidemiology.<br />

Expression <strong>of</strong> antimicrobial activity in food and clinical<br />

Listeria monocy<strong>to</strong>genes isolates<br />

Barbosa, J., Ferreira, V., Borges, S., Azevedo, I., Magalhães, R., San<strong>to</strong>s, I, Almeida, G.<br />

and Teixeira, P.<br />

Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Portugal<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive facultative intracellular pathogen capable<br />

<strong>of</strong> causing fatal infections in risk groups, such as pregnant women, elderly and<br />

immunocompromised individuals. Little information is available on the antimicrobial<br />

susceptibility <strong>of</strong> L. monocy<strong>to</strong>genes, particularly <strong>of</strong> strains isolated from food<br />

and production environments. Minimum inhibi<strong>to</strong>ry concentrations (MIC’s) <strong>of</strong><br />

ampicillin, penicillin G, chloramphenicol, erythromycin, tetracycline and vancomycin<br />

were assessed for 370 and 118 food and clinical L. monocy<strong>to</strong>genes isolates,<br />

respectively. All the clinical isolates were susceptible <strong>to</strong> all the antibiotics tested.<br />

Amongst food isolates, 0.5 % were resistant <strong>to</strong> tetracycline and 4.1 % and 2.7 % were<br />

intermediary and resistant, respectively, <strong>to</strong> erythromycin. Only one isolate demonstrated<br />

a resistance pr<strong>of</strong>ile <strong>to</strong> tetracycline, an antibiotic largely used in veterinary<br />

practice, but the resistance <strong>of</strong> nine isolates <strong>to</strong> erythromycin is a reason for concern,<br />

since erythromycin is used <strong>to</strong> treat pregnant women diagnosed with listeriosis or as<br />

a drug <strong>of</strong> second choice in cases <strong>of</strong> ampicillin allergy. Although in vitro antibiotic<br />

susceptibility testing does not always reflect the in vivo situation, results demonstrated<br />

that some <strong>of</strong> the isolates investigated are resistant <strong>to</strong> antibiotics <strong>of</strong> clinical<br />

importance commonly used <strong>to</strong> treat listeriosis. Listeriosis is usually an easily treatable<br />

disease and no reports have been presented on antibiotic resistant isolates<br />

from disease cases. However, it is an important area <strong>to</strong> keep under surveillance.<br />

Whilst the present study does not in a scientific sense present de novo science, it<br />

contains very valuable and important data. Only by keeping an on-going track <strong>of</strong><br />

the resistance patterns can these be moni<strong>to</strong>red for unexpected changes.<br />

R E F E R E N C E<br />

A / P<br />

29<br />

R E F E R E N C E<br />

A / P<br />

30


R E F E R E N C E<br />

A / P<br />

31<br />

R E F E R E N C E<br />

A / P<br />

32<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

The Listeria monocy<strong>to</strong>genes sigma B and sigma H regulons overlap, but<br />

only sigma B appears <strong>to</strong> be important for survival <strong>of</strong> acid, alkaline and<br />

oxidative stress<br />

Chaturongakul, S. 1, Raengpradub, S. 2, Wiedmann, M. 3 and Boor, K. J. 3<br />

1. Mahidol University, Thailand<br />

2. Silliker, Inc., USA<br />

3. Cornell University, USA<br />

Previous transcrip<strong>to</strong>mic analyses in Listeria monocy<strong>to</strong>genes using whole genome microarrays<br />

and Hidden Markov Model (HMM) based promoter searches identified<br />

many genes that were controlled by more than one transcriptional regula<strong>to</strong>r (e.g., σ B,<br />

σ H, σ L, PrfA, HrcA or CtsR). The regulons <strong>of</strong> the two stationary phase sigma fac<strong>to</strong>rs,<br />

σ B and σ H shared the largest number <strong>of</strong> genes. To further investigate the regula<strong>to</strong>ry<br />

features shared between σ B and σ H, whole genome microarrays were performed and<br />

transcript pr<strong>of</strong>iles for ∆sigB, ∆sigH, and ∆sigB∆sigH strains were compared <strong>to</strong> that <strong>of</strong><br />

the parent strain 10403S. In the present study, using cut-<strong>of</strong>f criteria <strong>of</strong> a 1.5 fold difference<br />

at p < 0.05, we identified 301 σ B-dependent genes, 82 s H-dependent genes,<br />

and a <strong>to</strong>tal <strong>of</strong> 61 genes with expression pr<strong>of</strong>iles that suggest regula<strong>to</strong>ry interactions<br />

between σ B and σ H. The 61 co-regulated genes grouped in<strong>to</strong> 4 different quality<br />

threshold (QT) clusters: Cluster 1 is comprised <strong>of</strong> 19 genes that are positively regulated<br />

by both σ B and σ H (e.g., mRNA levels for cspD, encoding a cold shock protein<br />

and for rpoD, encoding an RNA polymerase sigma fac<strong>to</strong>r, were significantly lower in<br />

the ∆sigB∆sigH strain than in the other strains), Cluster 2 contains 17 genes that are<br />

negatively regulated by both σ B and σ H (e.g., mRNA levels for genes encoding ribosomal<br />

proteins (rpsR and rplJ), RNA polymerase beta subunit (rpoB), and a sigma fac<strong>to</strong>r<br />

(sigC) were significantly higher in the ∆sigB∆sigH strain than in the other strains),<br />

Cluster 3 contains 11 genes with reduced expression in both the ∆sigB and ∆sigH<br />

strains, but with transcript levels that are the same in both the ∆sigB∆sigH and wild<br />

type strain (e.g., ATP-dependent Clp protease clpP), and Cluster 4 has 14 genes that<br />

are positively regulated in the ∆sigB strain but that are negatively regulated in the<br />

∆sigB∆sigH strain or vice versa (e.g., mreD, which contributes <strong>to</strong> cell shape and rpsD<br />

and rplK, which encode ribosomal proteins). In addition <strong>to</strong> transcript pr<strong>of</strong>iling,<br />

10403S and the 3 otherwise isogenic mutant strains were also phenotypically characterized.<br />

After exposure <strong>to</strong> acid (1 h at final pH <strong>of</strong> 2.5), alkaline (1 h at final pH <strong>of</strong> 12), or<br />

oxidative stresses (15 min at 13 mM cumene hydroperoxide final concentration), the<br />

phenotype <strong>of</strong> the ∆sigB∆sigH strain was identical <strong>to</strong> that <strong>of</strong> the ∆sigB strain, i.e., with<br />

cell numbers that were significantly reduced (CFU/ml) when compared <strong>to</strong> those <strong>of</strong><br />

the parent strain. This latter finding further verifies the central role <strong>of</strong> σ B in stress response,<br />

and suggests that σ H is less important for survival <strong>of</strong> these stresses.<br />

Virulence gene expression in Listeria monocy<strong>to</strong>genes strains isolated<br />

from different sources<br />

Alessandria, V., Rantsiou, K. and Cocolin, L.<br />

University <strong>of</strong> Turin, Italy<br />

Listeria monocy<strong>to</strong>genes is an important food-borne pathogen that has the capacity <strong>to</strong><br />

cause severe infections. Ubiqui<strong>to</strong>us microorganism, it is commonly isolated from<br />

foods <strong>of</strong> animal origin, mainly meat and milk products. However, due <strong>to</strong> its capacity <strong>to</strong><br />

develop at refrigeration temperatures, human listeriosis outbreaks are most <strong>of</strong>ten associated<br />

with ready-<strong>to</strong>-eat food products that are consumed without prior cooking.<br />

The incidence <strong>of</strong> the disease depends on different fac<strong>to</strong>rs including the infective dose<br />

and the immunity conditions <strong>of</strong> the host. The present work focuses on the expression<br />

analysis <strong>of</strong> four virulence genes (sigB, plcA, hly and iap) in 11 different strains <strong>of</strong> L.<br />

monocy<strong>to</strong>genes and, more in detail, 3 collection strains (EGDe, NCTC10527, SCOTT<br />

A), 7 isolated from food matrices (4 <strong>of</strong> which isolated from meat products and three<br />

from diary products) and 1 isolated from humans. In the first step, the expression<br />

analysis was performed in vitro, culturing each strain in BHI (Brain Heart Infusion)<br />

medium, in order <strong>to</strong> identify possible differences in the expression level for the different<br />

strains. The combined use <strong>of</strong> reverse transcription and quantitative PCR (qRT-<br />

PCR) was used <strong>to</strong> evaluate gene expression. As a second step, we wanted <strong>to</strong> evaluate<br />

the trend <strong>of</strong> expression <strong>of</strong> the genes in food. Analyses were performed inoculating L.<br />

monocy<strong>to</strong>genes in milk at different conditions and in particular at two temperatures<br />

(4 and 12 °C) for different times (24 and 48 hours). Significant expression differences<br />

emerged for the different genes, without showing any significant association between<br />

the expression <strong>of</strong> the genes and the origin <strong>of</strong> the different strains.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Differentiation <strong>of</strong> Listeria monocy<strong>to</strong>genes, Listeria innocua and<br />

Listeria marthii, a novel Listeria species isolated from the natural<br />

environment, Finger Lakes National Forest<br />

Graves, L. M. 1, Helsel, L. O. 1, Steigerwalt, A. G. 1, Morey, R. E. 1, Daneshvar, M. I. 1,<br />

Ro<strong>of</strong>, S. E., 2 Orsi, R. H. 2, Fortes, E. D. 2, Milillo, S. R. 2, den Bakker, H. C. 2, Wiedmann, M. 2,<br />

Swaminathan, B. 1 and Sauders, B. D. 3<br />

1. Centers for Disease Control and Prevention, USA<br />

2. Cornell University, USA<br />

3. New York State Department <strong>of</strong> Agriculture and Markets Food Labora<strong>to</strong>ry Division, USA<br />

Four isolates (FSL S4-120 T, FSL S4-696, FSL S4-710, and FSL S4-965) <strong>of</strong> Grampositive,<br />

motile, facultatively anaerobic, non-sporeforming bacilli that were phenotypically<br />

similar <strong>to</strong> Listeria spp. were isolated from soil, standing water, and flowing<br />

water samples obtained from the natural environment in the Finger Lakes<br />

National Forest, New York, USA. The four isolates were closely related <strong>to</strong> one another<br />

and were determined the same species by whole genome DNA-DNA hybridization<br />

studies, but sufficiently different from L. monocy<strong>to</strong>genes and L. innocua<br />

by DNA-DNA hybridization <strong>to</strong> form a separate species. 16S ribosomal RNA sequence<br />

analysis confirmed their close phylogenetic relatedness <strong>to</strong> L. monocy<strong>to</strong>genes<br />

and L. innocua and more distant relatedness <strong>to</strong> L. welshimeri, L. seeligeri, L. ivanovii,<br />

and L. grayi. Phylogenetic analysis <strong>of</strong> partial sequences for sigB, gap, and prs<br />

showed that these isolates form a well-supported sistergroup <strong>to</strong> L. monocy<strong>to</strong>genes.<br />

The four isolates showed > 82 % relatedness at 55 °C and > 76 % relatedness at 70 °C<br />

with 0.0 – 0.5 % divergence. Labeled DNA from the type strain (FSL S4-120 T)<br />

showed an average <strong>of</strong> 73 % relatedness <strong>to</strong> three L. monocy<strong>to</strong>genes strains and one L.<br />

innocua strain (range, 69 – 75 %) in reactions at 55 °C, the divergence in related DNA<br />

sequences was between 7.5 % and 9.5 %. In reactions at 70 °C, FSL S4-120 T DNA<br />

showed an average <strong>of</strong> 45 % relatedness <strong>to</strong> the three L. monocy<strong>to</strong>genes strains and<br />

one L. innocua strain (range, 30 – 53 %). The isolates yielded positive reactions in<br />

the AccuProbe® test that is purported specific for L. monocy<strong>to</strong>genes, did not ferment<br />

L-rhamnose, were non-hemolytic on blood agar media, and did not contain a<br />

homologue <strong>of</strong> the L. monocy<strong>to</strong>genes virulence gene island. A new species L. marthii<br />

is described which phenotypically and genotypically is distinct from all other Listeria<br />

species including its closest neighbors L. monocy<strong>to</strong>genes and L. innocua.<br />

Differed roles <strong>of</strong> L,D-carboxypeptidases encoded by lmo0028<br />

and lmo1638 genes.<br />

Yurov, D.*, Varfolomeev, A., Kaminskaya, A. and Ermolaeva, S.<br />

Gamaleya Institute <strong>of</strong> Epidemiology and Microbiology, Moscow, Russia<br />

L,D-carboxypeptidase hydrolyses a bound between m-A 2 pm and C-terminal Dalanine<br />

in murein. Listeria monocy<strong>to</strong>genes carried two genes, lmo0028 and<br />

lmo1638 that encode L,D-carboxypeptidases. The lmo0028 promoter has a recognition<br />

site for the virulence regula<strong>to</strong>r PrfA. Mutant L. monocy<strong>to</strong>genes strains<br />

GIM0028 and GIM1638 were obtained by site-specific insertions in<strong>to</strong> the chromosome<br />

<strong>of</strong> the wild type EGDe strain. Morphology was characterized with light<br />

microscopy. Septation was studied with fluorescent vancomycin. Genome copy<br />

numbers were determined with quantitative PCR. L. monocy<strong>to</strong>genes virulence was<br />

studied on BALB/c mice and human colon carcinoma HT29 cells. The mutation<br />

in lmo0028 caused dispersion in cell lengths: the length/width ratio ranged between<br />

1.5 and 15 for GIM0028 and 1 and 4.4 for EGDe. The insertion in lmo1638<br />

caused shortening in cells. The average length/width ratios were 3.75, 1.2 and 3.0<br />

for GIM0028, GIM1638 and EGDe, respectively. The longish GIM0028 cells were<br />

defected in septum formation, GIM1638 were not. GIM0028 demonstrated two<br />

fold decreasing in invasion efficiency. Doubling times in HT29 cells determined by<br />

plating were 51 minutes and 70 minutes for EGDe and GIM0028, respectively.<br />

However, evaluation <strong>of</strong> intracellular bacterial genome numbers with qPCR did not<br />

reveal a difference between the strains. GIM0028 bacteria increased their length<br />

upon intracellular growth: average length/width ratio was about 4.8 for invading<br />

bacteria and 9.2 after 6 h <strong>of</strong> intracellular growth while for the wild type bacteria it<br />

was 3.5 (p < 0.005). GIM0028 demonstrated reduced virulence for BALB/c mice<br />

with LD 50 <strong>of</strong> 1x10 6 and 1.5x10 4 CFU/mouse for GIM008 and EGDe strains, respectively.<br />

The Lmo0028 L,D-carboxypeptidase is involved in septum formation<br />

while Lmo1638 seems <strong>to</strong> be involved in cell elongation. A mutation in lmo0028<br />

decreases L. monocy<strong>to</strong>genes virulence.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

A / P<br />

33<br />

R E F E R E N C E<br />

A / P<br />

34


R E F E R E N C E<br />

A / P<br />

35<br />

R E F E R E N C E<br />

A / P<br />

36<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Antimicrobial susceptibilities <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

from retail beef, pork and poultry in Japan<br />

Ida, M., Shimojima, Y., Kaneko, S., Higuchi, Y., Nakama, A. and Kai, A.<br />

Tokyo Metropolitan Institute <strong>of</strong> Public Health, Japan<br />

Meat and meat products are the major source <strong>of</strong> food-borne infections and the<br />

most important link between food-producing animals and humans. The aim <strong>of</strong><br />

this study is evaluate antimicrobial susceptibilities <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

from meat in Japan. A <strong>to</strong>tal <strong>of</strong> 125 strains isolated from beef (n = 10), pork<br />

(n = 49) and poultry (n = 66) between 2001 and 2009 in Tokyo area were examined.<br />

Serotypes <strong>of</strong> tested strains were 1/2a (n = 46), 1/2b (n = 26), 1/2c (n = 15), 4b (n = 26)<br />

and others (n = 12, including untypable). Ampicillin, penicillin G, gentamicin,<br />

kanamycin, erythromycin, oxytetracycline, norfloxacin and chloramphenicol were<br />

included in the study. The minimum inhibi<strong>to</strong>ry concentrations <strong>of</strong> antimicrobial<br />

agents were determined by the agar dilution method according <strong>to</strong> the guidelines <strong>of</strong><br />

Clinical Labora<strong>to</strong>ry Standards Institutes (CLSI). CLSI breakpoint guidelines for L.<br />

monocy<strong>to</strong>genes and other Gram-positive bacteria were used for data analysis. Enterococcus<br />

faecalis ATCC 29212 and Escherichia coli ATCC 25922 were used as<br />

quality control strains. All isolates were susceptible <strong>to</strong> antimicrobials except for<br />

oxytetracycline. Three strains, isolated from domestic pork, domestic poultry and<br />

beef imported from Australia were resistant (> 32µg/ml). All <strong>of</strong> them were isolated<br />

in 2001 and belonging <strong>to</strong> serotype 1/2c. This study showed that antimicrobial resistance<br />

is not highly prevalent in L. monocy<strong>to</strong>genes in meat in Japan.<br />

Genetic basis <strong>of</strong> two low pathogenic L. monocy<strong>to</strong>genes strains<br />

with apparent phospholipase C activity<br />

Jiang, L., Bai, F., Chen, J., and Fang, W.<br />

Zhejiang University Institute <strong>of</strong> Preventive Veterinary Medicine, Hangzhou, China<br />

Two Listeria monocy<strong>to</strong>genes isolates showing apparent in vitro phospholipase C<br />

activity were characterized for their genetic basis and pathogenecity. Sequencing<br />

analysis found two major mutations in plcB at position 1 [from A (ATG) <strong>to</strong> G<br />

(GTG)] and position -26 [from C (ACG) <strong>to</strong> T(ATG)], resulting in 9-aa extension <strong>of</strong><br />

the enzyme at the N-terminus. The strains were <strong>of</strong> serovar 4a, had low pathogenecity<br />

with LD 50 at log 10 CFU 8.21-8.35 in BALB/c mice, and presented no microscopically<br />

visible plaques on the L929 cell monolayers. Site-directed mutation<br />

<strong>of</strong> these sites <strong>to</strong> the plcB version <strong>of</strong> strain 10403S abolished the enzyme activity, increased<br />

virulence by one log (from 8.12 <strong>to</strong> 7.07), but did not have significant effect<br />

on plaquing. This abolishment did not seem <strong>to</strong> be related <strong>to</strong> presence <strong>of</strong> G145S in<br />

the central regula<strong>to</strong>r prfA, which was found <strong>to</strong> induce overexpression <strong>of</strong> virulence<br />

genes independent <strong>of</strong> environmental conditions in other strains. However, this<br />

does not mean that prfA was not involved in regulation <strong>of</strong> its expression because<br />

transposon mutagenesis identified two mutant strains lacking apparent phospholipase<br />

activity due <strong>to</strong> disruption <strong>of</strong> the prfA by insertions at nt364 and nt828. T<strong>here</strong><br />

are two additional changes <strong>of</strong> T165A and K197N in prfA, as compared with other L.<br />

monocy<strong>to</strong>genes serovar 1/2a and 4b strains. Further work is under way <strong>to</strong> examine<br />

if apparent activity <strong>of</strong> the novel phospholipase in these strains results from easy<br />

access <strong>to</strong> zinc metalloprotease encoded by mpl for maturation, or if amino acid<br />

substitutions <strong>of</strong> T165A and K197N other than G145S in prfA also contribute <strong>to</strong> regulation<br />

<strong>of</strong> downstream virulence genes. Using one <strong>of</strong> the strains as a model organism,<br />

we are testing how phospholipase C and listerolysin interact in determining<br />

its pathogenecity and if t<strong>here</strong> are other genes involved in their low pathogenecity<br />

by genome wide scanning using Solexa sequencing technology.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Molecular genotyping and antimicrobial resistance<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes from foods and the environment<br />

Parisi, A. 1, Miccolupo, A. 1, Fraccalvieri, R. 1, La<strong>to</strong>rre, L. 1, Normanno, G. 2 and Santagada, G. 1<br />

1. Experimental Zooprophylactic Institute Apulia and Basilicata, Italy<br />

2. University <strong>of</strong> Bari, Faculty <strong>of</strong> Veterinary Medicine, Italy<br />

Listeria monocy<strong>to</strong>genes is generally susceptible <strong>to</strong> a wide range <strong>of</strong> antibiotics, however<br />

an increasing number <strong>of</strong> strains resistant <strong>to</strong> one or more antibiotics have been<br />

reported. Objectives <strong>of</strong> the present study were <strong>to</strong> evaluate molecular genotyping<br />

and antimicrobial susceptibility in L. monocy<strong>to</strong>genes isolates from foods and the<br />

environment. Over all a <strong>to</strong>tal <strong>of</strong> 100 L. monocy<strong>to</strong>genes isolates from foods (n = 74)<br />

and the environment (n = 26) were subjected <strong>to</strong> molecular genotyping using Multi-<br />

Locus Sequence Typing as previously described, moreover the susceptibility <strong>to</strong> 21<br />

antimicrobials was performed using Gram-positive GPN4F kit (TREK Diagnostic<br />

System). All the isolates resulted sensitive <strong>to</strong> ampicillin, chloramphenicol, gentamycin,<br />

penicillin G, rifampicin, strep<strong>to</strong>mycin, sulfamethoxazole trimethoprim.<br />

Antimicrobial resistance was recorded for oxacyllin (24 %), and tetracycline (1 %)<br />

w<strong>here</strong>as some isolates resulted moderately sensitive <strong>to</strong>ward clindamycin (19 %),<br />

quinupristin/dalfopristin (2 %) and cipr<strong>of</strong>loxacin (1 %). The isolates belonged <strong>to</strong><br />

six different serotypes: 1/2a (43 %), 1/2c (24 %), 1/2b (15 %), 4b/4e (12 %), 3a (4 %)<br />

and 3b (2 %). MLST identified 50 different Sequence Types (ST) most <strong>of</strong> which<br />

represented by a single isolate. ST9 (21 %), ST121 (12 %), ST3 (5 %) and ST199 (5 %)<br />

were the most prevalent STs. Moreover MLST allowed a good discrimination <strong>of</strong><br />

the two main genetic lineage <strong>of</strong> L. monocy<strong>to</strong>genes. Our results confirm the low<br />

prevalence <strong>of</strong> food and environmental isolates resistant <strong>to</strong> antimicrobials. Interestingly,<br />

a good correlation between resistance <strong>to</strong> oxacillin and genetic lineage I<br />

was observed (Yates corrected = 35,46; p < 0.005); this is in agreement with the<br />

clonal organization <strong>of</strong> L. monocy<strong>to</strong>genes population. The severity <strong>of</strong> L. monocy<strong>to</strong>genes<br />

infections and the reported increasing <strong>of</strong> antimicrobial resistance make necessary<br />

the continuous moni<strong>to</strong>ring <strong>of</strong> this trend although the recorded data do not<br />

appear <strong>to</strong> be alarming. The comparison <strong>of</strong> phenotypic and genetic characters allow<br />

<strong>to</strong> validate the evolutionary models speculated for this microbial species.<br />

Virulence transcrip<strong>to</strong>me analysis <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

by application <strong>of</strong> microarrays in vitro and in situ<br />

Rantsiou, K., Alessandria, V. and Cocolin, L.<br />

University <strong>of</strong> Turin, Italy<br />

With the complete genome sequence available for Listeria monocy<strong>to</strong>genes as well<br />

as other Listeria spp., it is nowadays possible <strong>to</strong> use <strong>to</strong>ols that allow global analysis<br />

<strong>of</strong> the biology <strong>of</strong> this pathogenic microorganism. Microarrays give the possibility <strong>to</strong><br />

study concurrently a large number <strong>of</strong> genes and deduce information regarding<br />

their expression. This approach accelerates significantly the rhythm by which new<br />

information is generated. In this way, the effect <strong>of</strong> different conditions on the expression<br />

<strong>of</strong> a series <strong>of</strong> genes and as a consequence on the physiology <strong>of</strong> a microorganism<br />

can be investigated. Within the 6 th FP project Pathogen Combat, it was<br />

developed a microarray including 72 genes <strong>of</strong> L. monocy<strong>to</strong>genes, encoding for virulence,<br />

adhesion and stress response genes. In this study we sought <strong>to</strong> apply this<br />

array <strong>to</strong> study the effect <strong>of</strong> different environmental conditions on the expression <strong>of</strong><br />

virulence genes. First, different strains <strong>of</strong> L. monocy<strong>to</strong>genes were grown in vitro, at<br />

different temperatures, pH and salt concentrations. Then, L. monocy<strong>to</strong>genes was<br />

artificially inoculated in different food matrices, <strong>to</strong> simulate the conditions <strong>of</strong> ‘virulence’<br />

<strong>of</strong> this microorganism at time <strong>of</strong> consumption. Global analysis <strong>of</strong> the expression<br />

<strong>of</strong> virulence genes in different L. monocy<strong>to</strong>genes strains will lead <strong>to</strong> a better<br />

undestanding <strong>of</strong> the physiology <strong>of</strong> this microorganism and eventually <strong>to</strong> the<br />

prediction <strong>of</strong> its potential <strong>to</strong> cause disease <strong>to</strong> humans.<br />

R E F E R E N C E<br />

A / P<br />

37<br />

R E F E R E N C E<br />

A / P<br />

38


R E F E R E N C E<br />

A / P<br />

39<br />

R E F E R E N C E<br />

A / P<br />

40<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Effects <strong>of</strong> growth conditions on surface properties <strong>of</strong> Listeria;<br />

a proposed role for AI-2<br />

Wong, H. T. L., Nwaiwu, O.* and Rees, C. E. D.<br />

School <strong>of</strong> Biosciences, University <strong>of</strong> Nottingham, UK<br />

Listeria monocy<strong>to</strong>genes is recognized as a particular problem in fac<strong>to</strong>ry environments<br />

due <strong>to</strong> its ability <strong>to</strong> colonize surfaces and then become persistent by adapting<br />

<strong>to</strong> low nutrient and low temperature conditions. We have been investigating<br />

the effect <strong>of</strong> low nutrient conditions on the ability <strong>of</strong> Listeria <strong>to</strong> attach <strong>to</strong> surfaces.<br />

Cultures were prepared using a minimal medium (D10), a rich medium (BHI) and<br />

also the D10 medium supplemented with duck meat extract. Growth in the BHI<br />

medium was fastest, but significant increases in growth were seen when D10 was<br />

supplemented with 10 % duck meat extract. One feature <strong>of</strong> cells that become persistent<br />

in fac<strong>to</strong>ry environments is their ability <strong>to</strong> form bi<strong>of</strong>ilms that are resistant <strong>to</strong><br />

cleaning and disinfection. We have found that the surface hydrophobicity (measured<br />

using Microbial Attachment <strong>to</strong> Solvents (MATS) assays) alters significantly<br />

when the organism is grown under low nutrient conditions, with cells becoming<br />

more hydrophobic. Also these cells have a greater tendency <strong>to</strong> flocculate in liquid<br />

culture. This affect is seen when cells are grown above 30 °C and t<strong>here</strong>fore is not<br />

related <strong>to</strong> the synthesis <strong>of</strong> flagellae. The ability <strong>to</strong> form bi<strong>of</strong>ilms in Listeria is<br />

known <strong>to</strong> be affected by the synthesis <strong>of</strong> a functional au<strong>to</strong>inducer 2 (AI-2)-like<br />

signal, and that an intact luxS gene is associated with inhibition <strong>of</strong> attachment and<br />

bi<strong>of</strong>ilm formation. Using a Vibrio harveyi reporter strain we have found that when<br />

cells are grown in minimal media, the amount <strong>of</strong> detectable AI-2 is reduced. As<br />

AI-2 has a role in the down-regulation <strong>of</strong> components required for attachment<br />

and bi<strong>of</strong>ilm formation, this result is consistent with our observation <strong>of</strong> changes<br />

in the surface properties <strong>of</strong> Listeria under low nutrient conditions that are more<br />

likely <strong>to</strong> promote surface attachment.<br />

* Participation Supported by IUFoST.<br />

Stress behaviour <strong>of</strong> a Listeria monocy<strong>to</strong>genes 568 Lmo1634<br />

transposon mutant<br />

Truelstrup-Hansen, L. 1, Holman, D. B. 1 and Ells, T. C. 2<br />

1. Dalhousie University, Canada<br />

2. Agriculture and Agri-Food Canada<br />

Listeria monocy<strong>to</strong>genes is an important psychrotrophic foodborne pathogenic bacterium<br />

with tendency <strong>to</strong> persist in food processing plants. Heat treatment is routinely<br />

used as a means <strong>to</strong> control pathogenic bacteria in food products. T<strong>here</strong>fore,<br />

the thermo<strong>to</strong>lerance <strong>of</strong> L. monocy<strong>to</strong>genes and mechanisms responsible are <strong>of</strong> interest.<br />

Previous research using transposon mutagenesis yielded several heat resistant<br />

mutants, including one, named 1B4, that was mapped <strong>to</strong> a putative alcohol/acetaldehyde<br />

dehydrogenase gene, lmo1634 or adh, also identified as Listeria adhesion<br />

protein. The objective <strong>of</strong> this study was <strong>to</strong> investigate the function <strong>of</strong> this gene<br />

in relation <strong>to</strong> a variety <strong>of</strong> environment stresses and in bi<strong>of</strong>ilm formation. The transposon<br />

insertional mutant 1B4 demonstrated significantly greater <strong>to</strong>lerance <strong>to</strong> acid<br />

(pH 2.7), 20 % NaCl, and heat treatment <strong>of</strong> 52 °C than the wild-type (WT) Lm568<br />

strain. Attachment <strong>of</strong> 1B4 (7.06 log 10 CFU/cm 2) <strong>to</strong> stainless steel coupons was not<br />

significantly (p > 0.05) different from the WT Lm568 (6.91 log 10 CFU/cm 2). Neither<br />

1B4 nor Lm568 exhibited greater survival after desiccation (43 % RH) at 20 °C for 7<br />

days. The alcohol dehydrogenase activity <strong>of</strong> Lm568 (0.197 U/mg) was found <strong>to</strong> be<br />

significantly (p > 0.05) greater than in 1B4 (0.0458 U/mg). 1B4 was complemented<br />

in trans with the WT lmo1634 gene and the successful transcription <strong>of</strong> lmo1634 in<br />

the complemented strain was confirmed by RT-PCR. In addition, no polar effect on<br />

the genes immediately downstream <strong>of</strong> lmo1634 in 1B4, i.e., lmo1635, lmo1636, or<br />

lmo1637, was observed. The trans complementation <strong>of</strong> 1B4, however, did not completely<br />

res<strong>to</strong>re the WT phenotype. Attempts <strong>to</strong> create a deletion mutant in lmo1634<br />

ultimately proved unsuccessful. Despite these technical issues, it is clear that a mutation<br />

in lmo1634 yields a multi-stress resistant phenotype.


AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Investigation <strong>of</strong> the conditions that trigger the activation <strong>of</strong> the alternative<br />

sigma fac<strong>to</strong>r σB in Listeria monocy<strong>to</strong>genes<br />

Utratna, M., Shaw, I. and O’Byrne, C.<br />

National University <strong>of</strong> Ireland, Galway, Ireland<br />

Listeria monocy<strong>to</strong>genes has evolved multiple strategies <strong>to</strong> overcome diverse stress<br />

conditions and it is able <strong>to</strong> occupy extreme environmental niches. As a facultative<br />

intracellular foodborne pathogen L. monocy<strong>to</strong>genes successfully invades host cells,<br />

surviving in gastric tract before the final colonization <strong>of</strong> the target organs. The alternative<br />

sigma fac<strong>to</strong>r (σ B) plays an essential role in the stress <strong>to</strong>lerance and virulence<br />

<strong>of</strong> L. monocy<strong>to</strong>genes. Many components <strong>of</strong> the σ B regulon have been identified<br />

but the mechanisms that regulate σ B are still unknown. The current model<br />

for σ B regulation, based on work done in Bacillus subtilis, suggests that the activity<br />

<strong>of</strong> σ B is independent <strong>of</strong> the levels <strong>of</strong> σ B in the cell. The expression <strong>of</strong> OpuCA has<br />

been shown <strong>to</strong> be controlled by σ B. T<strong>here</strong>fore the levels <strong>of</strong> this protein can act as an<br />

indirect reporter <strong>of</strong> the activity <strong>of</strong> σ B in L. monocy<strong>to</strong>genes. Polyclonal antibodies<br />

against OpuCA were developed in chickens and used in Western blotting. Based<br />

on this reporter system it was shown that σ B is more active in stationary phase<br />

when compared <strong>to</strong> exponential phase (OD 600 =0.6). σ B activity under conditions<br />

encountered in the human gut was also determined. Levels <strong>of</strong> OpuCA increase<br />

gradually in the range <strong>of</strong> salt up <strong>to</strong> 0.9M NaCl and when the pH is reduced from pH<br />

7.2 <strong>to</strong> pH 5.0. Higher σ B activity was also detected under limited oxygen conditions<br />

and at 4 °C. However, no change in the levels <strong>of</strong> OpuCA was observed in the<br />

presence <strong>of</strong> low concentrations <strong>of</strong> bile salts up <strong>to</strong> 5mM. Additionally an attempt<br />

was made <strong>to</strong> develop a transcriptional reporter fusion system by cloning the σ B<br />

promoter from strongly σ B-dependent gene, lmo2230, in<strong>to</strong> the reporter vec<strong>to</strong>r<br />

pTCV-lac allowing the measurement <strong>of</strong> β-galasc<strong>to</strong>sidase activity under various<br />

conditions. This genetic system will also be a useful <strong>to</strong>ol in moni<strong>to</strong>ring the activity<br />

<strong>of</strong> σ B in L. monocy<strong>to</strong>genes.<br />

Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes 1/2a, 1/2b, 1/2c and 4b<br />

by Amplified Fragment Length Polymorphism and evaluation <strong>of</strong> their<br />

geographical distributions in Portugal<br />

Maia, C. H. 1, Goulão, M. M. 2, San<strong>to</strong>s, M. I. 1, Ferreira, M. A. S. S. 3 and Pintado, C. M. B. S. 2<br />

1. Institu<strong>to</strong> Nacional de Saúde Dou<strong>to</strong>r Ricardo Jorge, Portugal<br />

2. Escola Superior Agrária, Institu<strong>to</strong> Politécnico de Castelo Branco, Portugal<br />

3. Institu<strong>to</strong> Superior de Agronomia, Universidade Técnica de Lisboa, Portugal<br />

Listeriosis is a bacterial infection caused by the ingestion <strong>of</strong> foods contaminated<br />

with Listeria monocy<strong>to</strong>genes, that can affect humans, especially the group that includes<br />

pregnant women, newborn infants, the elderly and immunocompromised<br />

individuals. AFLP molecular subtyping, using EcoRI restriction enzyme, was used<br />

<strong>to</strong> do the differentiation between strains isolated from different Denomination <strong>of</strong><br />

Protected Origin regions <strong>of</strong> cheese production in Portugal and as well from clinical<br />

specimens. A set <strong>of</strong> 159 strains belonging <strong>to</strong> serotypes 1/2a, 1/2b, 1/2c and 4b isolated<br />

from samples <strong>of</strong> Serra da Estrela ewe’s cheese, São Jorge cow´s cheese, Serpa<br />

ewe´s cheese, Tolosa ewe´s cheese and Castelo Branco ewe´s cheese were analyzed.<br />

Evaluation <strong>of</strong> similarities between strains was obtained by analyses with BioNumerics<br />

s<strong>of</strong>tware by Applied Maths, Belgium. With AFLP molecular subtyping,<br />

nearly all the isolates were divided according <strong>to</strong> their serovar. T<strong>here</strong>fore it was<br />

possible <strong>to</strong> predict the serovar from the AFLP pattern found in most <strong>of</strong> the isolates.<br />

In the opposite direction, each one <strong>of</strong> the serovars was associated <strong>to</strong> several<br />

AFLP molecular types, which demonstrates the grater discrimina<strong>to</strong>ry capacity <strong>of</strong><br />

the AFLP subtyping. Similarities between pr<strong>of</strong>iles, based on band positions, were<br />

derived from the Dice correlation coefficient (S D ) with a maximum position <strong>to</strong>lerance<br />

<strong>of</strong> 1 %, 3 % or 4 %. Listeria monocy<strong>to</strong>genes strains were clustered by the technique<br />

<strong>of</strong> the UPGMA and a dendrogram was constructed <strong>to</strong> reflect the genetic distance<br />

between them.<br />

R E F E R E N C E<br />

A / P<br />

41<br />

R E F E R E N C E<br />

A / P<br />

42


R E F E R E N C E<br />

A / P<br />

43<br />

R E F E R E N C E<br />

A / P<br />

44<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Global analysis <strong>of</strong> the Listeria monocy<strong>to</strong>genes surface proteins<br />

<strong>of</strong> the LPXTG family<br />

Botello-Morte, L. 1, Calvo, E. 2, Mariscotti, J. 1, D’Orazio, V. 1, García-del Portillo, F. 1<br />

and Pucciarelli, M. G. 1,3<br />

1. Departamen<strong>to</strong> de Biotecnología Microbiana, Centro Nacional de Biotecnología, Consejo Superior<br />

de Investigaciones Científicas (CSIC). Darwin 3, Madrid, Spain<br />

2. Unidad de Proteómica, Centro Nacional de Investigaciones Cardiovasculares (CNIC), Madrid, Spain<br />

3. Departamen<strong>to</strong> de Biología Molecular, Universidad Autónoma de Madrid. Campus de Can<strong>to</strong>blanco,<br />

Madrid, Spain<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive intracellular bacterial pathogen that<br />

causes serious systemic diseases in humans and animals. Comparative genomics<br />

revealed the presence in the Listeria genus <strong>of</strong> a large number <strong>of</strong> genes encoding<br />

surface proteins containing an LPXTG sorting motif. This signature makes these<br />

proteins recognized by sortases, the enzymatic machinery that anchors proteins<br />

covalently <strong>to</strong> the peptidoglycan. In the case <strong>of</strong> the L. monocy<strong>to</strong>genes EGD-e strain,<br />

a <strong>to</strong>tal <strong>of</strong> 41 genes encoding LPXTG proteins were discovered. Despite the relevance<br />

<strong>of</strong> surface proteins for communication with the environment and the host,<br />

the biological role <strong>of</strong> most members <strong>of</strong> this family remains unknown in L. monocy<strong>to</strong>genes.<br />

In a collaborative effort with other European groups that contributed <strong>to</strong><br />

generate mutants defective in specific LPXTG proteins, we accomplished a comprehensive<br />

proteomic analysis in peptidoglycan material purified from each <strong>of</strong><br />

these mutants. A <strong>to</strong>tal <strong>of</strong> 30 LPXTG mutants have been analysed <strong>to</strong> date. Proteomic<br />

analysis was also performed in peptidoglycan purified from intracellular<br />

wild-type bacteria upon infection <strong>of</strong> epithelial cells. A major finding <strong>of</strong> this global<br />

approach was the cross-regulation existing among certain LPXTG proteins. This<br />

phenomenon involved in most cases the down-regulation <strong>of</strong> concrete LPXTG proteins<br />

in response <strong>to</strong> the lack <strong>of</strong> a specific LPXTG protein. Interestingly, the inverse<br />

phenomenon, the up-regulation <strong>of</strong> a LPXTG protein in response <strong>to</strong> a deficiency<br />

in another LPXTG protein, was also observed. These proteomic data were confirmed<br />

with specific antibodies. Proteins subjected <strong>to</strong> these regula<strong>to</strong>ry processes,<br />

currently investigated in detail by our group, include Lmo0320 (Vip), Lmo0514,<br />

Lmo0610, Lmo0880, InlE, InlG, InlH, and Lmo2085.<br />

Antimicrobial susceptibility <strong>of</strong> Listeria monocy<strong>to</strong>genes strains<br />

derived from food and food-processing bakery plant<br />

Eusébio, C., Carneiro, L., San<strong>to</strong>s, I., Magalhães, R., Almeida, G., Silva, J. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

The susceptibility <strong>of</strong> 167 strains <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated from a bakery<br />

industry, from food and food-processing environment, <strong>to</strong> 11 antibiotics was determined<br />

by the standard agar dilution methodology. The tested antibiotics were:<br />

ampicillin, cipr<strong>of</strong>loxacin, chloramphenicol, erythromycin, gentamicin, nitr<strong>of</strong>uran<strong>to</strong>in,<br />

penicillin, rifampicin, strep<strong>to</strong>mycin, tetracycline and vancomycin; minimal<br />

inhibi<strong>to</strong>ry concentrations values were used <strong>to</strong> classify the strains in<strong>to</strong> sensitive,<br />

moderately resistance and resistant. All the tested isolates were found <strong>to</strong> be<br />

susceptible <strong>to</strong> ampicillin, chloramphenicol, gentamicin, nitr<strong>of</strong>uran<strong>to</strong>in, penicillin,<br />

rifampicin, tetracycline strep<strong>to</strong>mycin and vancomycin. In the case <strong>of</strong> erythromycin,<br />

54 isolates (32 %) were susceptible, 68 (41 %) displayed moderately resistance<br />

and 45 (27 %) were resistance. Concerning <strong>to</strong> cipr<strong>of</strong>loxacin, a moderate resistance<br />

was observed in 12 strains (7 %) against 155 strains (93 %) that were susceptible.<br />

Generally, this study showed that L. monocy<strong>to</strong>genes strains are susceptible <strong>to</strong> the<br />

antibiotics commonly used in the treatment <strong>of</strong> listeriosis. Concerning that antibiotic<br />

resistance in some L. monocy<strong>to</strong>genes strains has already been described, a continued<br />

study <strong>of</strong> emerging antimicrobial resistance is important <strong>to</strong> guarantee an<br />

effective treatment <strong>of</strong> human listeriosis.


A physiological study <strong>to</strong> purpose a new formal method<br />

<strong>to</strong> obtain L. monocy<strong>to</strong>genes cells adapted <strong>to</strong> BAC<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 00–46<br />

Saá Ibusquiza, P., Cabo, M. L., Herrera, J. J. R., Vázquez, D., Carrera, S. and Eiriz, E.<br />

Institu<strong>to</strong> de Investigaciones Marinas (Marine Research Institute), CSIC, Vigo, Galicia, España<br />

Maximum adaptation potential <strong>of</strong> Listeria monocy<strong>to</strong>genes (L. monocy<strong>to</strong>genes) cells<br />

<strong>to</strong> benzalkonium chloride (BAC) was classically checked by successive exposition<br />

<strong>of</strong> stationary cells <strong>to</strong> increasing sublethal BAC concentrations. However, this<br />

method is time consuming and do not ensure we reach the maximum level <strong>of</strong> adaptation<br />

<strong>of</strong> the strain. So, in the present work we develop a new method <strong>to</strong> obtain<br />

BAC-adapted cells <strong>of</strong> L. monocy<strong>to</strong>genes that is based in only one exposition <strong>of</strong><br />

L. monocy<strong>to</strong>genes exponential-phase cells <strong>to</strong> sublethal concentrations <strong>of</strong> BAC.<br />

Moreover, in order <strong>to</strong> predict the optimal conditions for achieving the maximum<br />

level <strong>of</strong> adaptation, a fac<strong>to</strong>rial design <strong>to</strong> formalize the effects <strong>of</strong> the initial concentration<br />

<strong>of</strong> cells and the concentration <strong>of</strong> BAC during the exposition was carried<br />

out. Obtained empirical model demonstrated a positive effect <strong>of</strong> inoculum and a<br />

positive interaction between both variables. However, a negative first order effect<br />

for the BAC indicated the convenience <strong>of</strong> using moderately high concentrations <strong>of</strong><br />

it during the expositions. As a consequence <strong>of</strong> applying this procedure, adaptation<br />

<strong>of</strong> L. monocy<strong>to</strong>genes 5873 <strong>to</strong> BAC was increased aproximately 4 times respecting <strong>to</strong><br />

the wild type strain in only 36 h. Finally, comparison between the expression pr<strong>of</strong>iles<br />

in the wild and the highest BAC-adapted variant was also carried out.<br />

Characterization <strong>of</strong> a RNA-helicase in the human pathogen<br />

Listeria monocy<strong>to</strong>genes<br />

Netterling, S. and Johansson, J.<br />

Department <strong>of</strong> Molecular Biology, Umeå University, Umeå, Sweden<br />

The intracellular pathogen Listeria monocy<strong>to</strong>genes grows in a wide range <strong>of</strong> milieus<br />

and shows a great acceptance <strong>to</strong> different environmental conditions, such as<br />

being able <strong>to</strong> grow in a span from -1 <strong>to</strong> 45 °C. Listeriosis can cause severe symp<strong>to</strong>ms<br />

such as meningitis, with high lethality rate. The ability <strong>to</strong> grow at low temperatures<br />

is one <strong>of</strong> the most interesting aspects <strong>of</strong> this pathogen. We have looked at<br />

RNA helicases; proteins that are able <strong>to</strong> unwind short dsRNA structures in an ATPdependent<br />

manner. Most mRNAs develop complex secondary and tertiary structures<br />

during the maturation process, these structures needs <strong>to</strong> be unwound before<br />

translation initiation. The stability <strong>of</strong> these complex structures is influenced<br />

by temperature, <strong>of</strong>ten becoming more rigid at lower temperatures t<strong>here</strong>by hindering<br />

e.g. translation, resulting in undesired cold-stabilized RNA structures. RNA<br />

helicases may function by relaxing local RNA-RNA interactions or by dissociating<br />

RNA from proteins. In L. monocy<strong>to</strong>genes RNA helicases have earlier been<br />

shown <strong>to</strong> be up-regulated during growth in a cold environment. RNA-helicases<br />

have been suggested, but not proven, <strong>to</strong> participate in small regula<strong>to</strong>ry RNA events.<br />

We have looked at a mutant lacking one DExH-box RNA helicase in L. monocy<strong>to</strong>genes.<br />

This DExH-box RNA helicase deletion strain displayed an inability <strong>to</strong> grow<br />

at 4 °C and had an impaired growth at temperatures below 37 °C. Motility was also<br />

affected in the deletion strain implying that it cannot spread in food sources as<br />

well as the wild-type strain. All phenotypes were rescued in a complemented<br />

strain. Our results indicate that RNA helicases is <strong>of</strong> high importance <strong>to</strong> the bacterium<br />

for growth at lower temperatures, and provides a link between RNA helicases<br />

and motility.<br />

R E F E R E N C E<br />

A / P<br />

45<br />

R E F E R E N C E<br />

A / P<br />

46


R E F E R E N C E<br />

B / P<br />

47<br />

R E F E R E N C E<br />

B / P<br />

48<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

A Listeria monocy<strong>to</strong>genes strain is still virulent despite non-functional<br />

major virulence genes: Optical Mapping shows a potential mechanism<br />

Roche, S. M. 1, Grépinet, O. 1, Corde, Y. 1, Teixeira, A. P. 1, Kerouan<strong>to</strong>n, A. 2,<br />

Témoin, S. 1, Mereghetti, L. 3, Brisabois, A. 2 and Velge, P. 1<br />

1. INRA, UR 1282 Infectiologie Animale et Santé Publique, F-37380 Nouzilly and IFR 136, Agents<br />

transmissibles et Infectiologie, France<br />

2. AFSSA LERQAP, Unité Caractérisation et Epidémiologie Bactérienne, F-94706 Maisons-Alfort,<br />

France<br />

3. Université François Rabelais de Tours, EA3854 “Bactéries et risque materno-foetal”, Tours, France<br />

and CHRU, Tours, France<br />

L. monocy<strong>to</strong>genes is a pathogenic species, but some strains exhibit low virulence. In<br />

previous studies, 24 naturally occurring low-virulence L. monocy<strong>to</strong>genes strains<br />

were identified using a method combining a plaque-forming (PF) assay with subcutaneous<br />

(SC) injection in<strong>to</strong> the left hind footpad <strong>of</strong> mice. Based on their phenotypic<br />

characteristics, these low-virulence strains have been assigned by cluster<br />

analysis <strong>to</strong> one <strong>of</strong> four groups. The 11 strains belonging <strong>to</strong> Group I exhibit a mutated<br />

PrfA w<strong>here</strong>as five out <strong>of</strong> the six strains belonging <strong>to</strong> Group III have causal<br />

mutations in plcA, inlA and inlB genes. New strains exhibiting the same PFGE specific<br />

pr<strong>of</strong>ile than the low-virulence Group III strains have been identified and characterized.<br />

All were low-virulence strains exhibiting the same mutations characteristic<br />

<strong>of</strong> the Group III strains, beside the A23 strain which has been<br />

characterized as a virulent strain in the mouse model. Interestingly, this strain<br />

exhibited the same mutations than the low-virulence Group III strains in the inlA,<br />

inlB, plcA genes and another one in the mpl gene. This led <strong>to</strong> expression <strong>of</strong> inactive<br />

InlB, PI-PLC and PC-PLC proteins and <strong>to</strong> a lack <strong>of</strong> InlA expression which are<br />

known <strong>to</strong> be important for virulence. Despite these mutations in these major virulence<br />

genes, the A23 L. monocy<strong>to</strong>genes strain was still virulent in vitro and in vivo.<br />

Analysis by Optical Mapping (Phylogene-France, OpGen-USA) <strong>of</strong> the A23 and two<br />

Group III strains in comparison with the EGDe genome showed specific fragments<br />

inserted in the genome <strong>of</strong> the Group III strains. Identification <strong>of</strong> the genes modified<br />

by the insertion could explain the virulence <strong>of</strong> the A23 strain compared <strong>to</strong> the<br />

Group III strains. Phylogenetic analysis <strong>of</strong> these strains could be fruitful <strong>to</strong> understand<br />

the evolution <strong>of</strong> virulence trait as well as plasticity <strong>of</strong> virulence genes.<br />

Protein expression <strong>of</strong> lineage I, II, and III Listeria monocy<strong>to</strong>genes<br />

strains in murine macrophages<br />

Donaldson, J. R., Nanduri, B., Pittman, J. R., Burgess, S. C. and Lawrence, M. L.<br />

Mississippi State University, USA<br />

In the current study, we compared the ability <strong>of</strong> Listeria monocy<strong>to</strong>genes strains from<br />

genetic lineages I, II, and III (serovar 1/2a strain EGD, serovar 4b strain F2365, and<br />

serovar 4a strain HCC23) <strong>to</strong> proliferate in the murine macrophage cell line J774.1.<br />

We found that the lineage III strain HCC23 was able <strong>to</strong> initiate an infection but<br />

could not establish prolonged infection within the macrophages. By contrast, strains<br />

EGD and F2365 proliferated within macrophages for at least 7 hr. We further characterized<br />

this interesting phenotypic difference by determining the protein expression<br />

pr<strong>of</strong>iles <strong>of</strong> these strains at 0 hr, 3 hr, and 5 hr post-infection using two dimensional<br />

liquid chroma<strong>to</strong>graphy coupled with electrospray ionization tandem mass<br />

spectrometry. We found that metabolic and cell wall associated proteins were expressed<br />

by all three strains at 3 hr post-infection. However, increased expression <strong>of</strong><br />

stress response and DNA repair proteins was detected in the lineage I and II strains<br />

at 5 hr post-infection. These proteins were not significantly increased in the lineage<br />

III strain at this time point. By comparing the protein expression patterns <strong>of</strong> these<br />

three L. monocy<strong>to</strong>genes strains during intracellular growth in macrophages, we were<br />

able <strong>to</strong> detect biological differences that may determine the ability <strong>of</strong> L. monocy<strong>to</strong>genes<br />

<strong>to</strong> survive and persist in macrophages.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in bovine mastitic milk samples:<br />

Possible source <strong>of</strong> food borne infection<br />

Kalorey, D. R. 1, Warke, S. 1, Kurkure, N. V. 1 and Barbuddhe, S. B. 2<br />

1. Nagpur Veterinary College, India<br />

2. ICAR Research Complex for Goa, India<br />

Listeria monocy<strong>to</strong>genes is a food borne pathogen responsible for listeriosis, disease<br />

characterized by meningitis, encephalitis, reproductive disorders and septicemia<br />

in human beings. One <strong>of</strong> the major source <strong>of</strong> Listeria is milk. A study was<br />

conducted <strong>to</strong> know the prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in mastitic milk from<br />

central India. Quarter milk samples (n = 1221) were collected from dairy cows and<br />

screened for sub clinical mastitis by California mastitis test (CMT). CMT positive<br />

samples were examined for the presence <strong>of</strong> L. monocy<strong>to</strong>genes following two step<br />

enrichment and plating on selective agar. Confirmation <strong>of</strong> isolates was based on<br />

biochemical tests, haemolysis on blood agar and CAMP test. The L. monocy<strong>to</strong>genes<br />

confirmed isolates were subjected <strong>to</strong> PCR assay for detection <strong>of</strong> virulence marker<br />

genes (hlyA, actA, iap and prf A). A <strong>to</strong>tal <strong>of</strong> 71 (5.81 %) milk samples harbored L.<br />

monocy<strong>to</strong>genes. The hlyA gene was detected in all the isolates. The hlyA and prfA<br />

genes were detected in 33 strains. The hlyA and actA were in seven strains. Eleven<br />

strains harbored hlyA, actA and iap genes. All isolated strains were subjected <strong>to</strong><br />

serotyping by PCR and revealed all <strong>to</strong> be <strong>of</strong> 4b, 4d and 4e serogroup. Excretion <strong>of</strong> L.<br />

monocy<strong>to</strong>genes in milk may contribute <strong>to</strong> food borne infection in human. Detection<br />

<strong>of</strong> 4b serogroup is <strong>of</strong> highly significance owing <strong>to</strong> its association with food borne<br />

listeriosis outbreaks.<br />

Agr-dependent peptide sensing in L. monocy<strong>to</strong>genes – Effects on bi<strong>of</strong>ilm<br />

formation, virulence and global gene expression<br />

Waidmann, M. S. 1*, Monk, I. R. 2, Auchter, M. 1, Preising, N. P. 1, Hill, C. 3 and Riedel, C. U. 1<br />

1. Institute <strong>of</strong> Microbiology and Biotechnology, University <strong>of</strong> Ulm, Germany<br />

2. Trinity College Dublin, Ireland<br />

3. University College Cork, Ireland<br />

Au<strong>to</strong>inducing peptides are signalling molecules used in population-dependent gene<br />

regulation <strong>of</strong> Gram-positive microorganisms. The best studied example <strong>of</strong> bacterial<br />

peptide sensing is the agr system <strong>of</strong> staphylococci. An operon with high structural<br />

and sequence similarity <strong>to</strong> the staphylococcal arg system was recently identified in<br />

the genome <strong>of</strong> L. monocy<strong>to</strong>genes EGDe. The listerial agr system is composed <strong>of</strong> a four<br />

gene operon with agrB involved in the proteolytic processing/export <strong>of</strong> the gene<br />

product <strong>of</strong> agrD, the posttranscriptionally modified signalling peptide, and<br />

agrA/agrC encoding a typical two-component system with histidine kinase (AgrC)<br />

and response regula<strong>to</strong>r (AgrA). Here, we present our recent advances in identifying<br />

and characterizing the native agr peptide <strong>of</strong> L. monocy<strong>to</strong>genens from spent cell culture<br />

supernatant and the effects <strong>of</strong> agr peptide sensing on bi<strong>of</strong>ilm formation, virulence<br />

and global gene expression. Furthermore, the mechanisms <strong>of</strong> agr-dependent<br />

gene regulation in L. monocy<strong>to</strong>genes were investigated. A ∆argD mutant showed a<br />

significant defect in bi<strong>of</strong>ilm formation. Invasion <strong>of</strong> EGDe ∆argD in<strong>to</strong> Caco-2 and<br />

Hep-2 cells was significantly reduced compared <strong>to</strong> EGDe wt. Moreover, the ∆argD<br />

mutant showed an unusual subcellular distribution after primary invasion. In line<br />

with these observations promoter activities <strong>of</strong> important virulence genes were reduced<br />

in the ∆argD mutant. Using bioluminescence in vivo imaging, a significant attenuation<br />

<strong>of</strong> EGDe ∆argD in a murine model <strong>of</strong> listeriosis could be shown. Moreover,<br />

microarray analysis revealed that expression <strong>of</strong> a large number <strong>of</strong> genes<br />

belonging <strong>to</strong> all functional categories was affected in EGDe ∆argD both in exponential<br />

and stationary growth phase indicating a global impact <strong>of</strong> agr-dependent peptide<br />

sensing on all aspects <strong>of</strong> physiology in L. monocy<strong>to</strong>genes.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

B / P<br />

49<br />

R E F E R E N C E<br />

B / P<br />

50


R E F E R E N C E<br />

B / P<br />

51<br />

R E F E R E N C E<br />

B / P<br />

52<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

agrD-deletion affects InlA- and InlB-regulation via<br />

a temperature-dependent and -independent mechanism<br />

Waidmann, M. S. 1*, Monk, I. R. 2 and Riedel, C. U. 1<br />

1. Institute <strong>of</strong> Microbiology and Biotechnology, University <strong>of</strong> Ulm, Germany<br />

2. University College Cork, Ireland<br />

Quorum sensing via the secretion <strong>of</strong> au<strong>to</strong>inducing peptides is a commonly used<br />

mechanism for gene regulation among Gram-positive bacteria. A well known example<br />

is the accessory gene regula<strong>to</strong>r (agr) in Staphylococcus aureus, which is<br />

known <strong>to</strong> affect virulence. Following the sequencing <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

EGDe genome and the identification <strong>of</strong> a homologous gene locus, suggestions <strong>of</strong> a<br />

regula<strong>to</strong>ry role in virulence arose. The first step <strong>of</strong> listerial pathogenic lifecycle is<br />

the entry in<strong>to</strong> a host cell, a process mediated by members <strong>of</strong> the Internalin family.<br />

T<strong>here</strong>fore, Internalin A and B are described as the main players for uptake by epithelial<br />

and endothelial cells. To testify the role <strong>of</strong> agr in regulation <strong>of</strong> these virulence<br />

fac<strong>to</strong>rs, we evaluated the invasive capability <strong>of</strong> an agrD deletion mutant,<br />

which lacks the putative au<strong>to</strong>inducing peptide, in<strong>to</strong> Caco-2 and HEp-2 cells. These<br />

assays showed that the Internalin A-dependent invasion <strong>of</strong> the mutant was significantly<br />

decreased in comparison <strong>to</strong> the wild type. The temperature chosen for<br />

growing pre-cultures did not influence this defect. By contrast, the effect <strong>of</strong> the<br />

agrD deletion for Internalin B-mediated invasion in<strong>to</strong> HEp-2 cells was only impaired<br />

in pre-cultures grown at 37 °C. Furthermore, the <strong>to</strong>tal numbers <strong>of</strong> invaded<br />

bacteria were much higher as compared <strong>to</strong> 30 °C. These results indicate a temperature-independent<br />

role <strong>of</strong> agr in the regulation <strong>of</strong> Internalin A and a temperaturedependent<br />

mechanism in case <strong>of</strong> Internalin B. This influence could occur via affecting<br />

either the expression itself or the activation <strong>of</strong> the identified regula<strong>to</strong>rs<br />

PrfA and σ B. According <strong>to</strong> these results agr seems <strong>to</strong> be an additional player for listerial<br />

virulence.<br />

* Participation Supported by IUFoST.<br />

Bovine cranial nerve Schwann cells express E-cadherin, a candidate<br />

key-player in the brainstem invasion <strong>of</strong> Listeria monocy<strong>to</strong>genes in cattle<br />

Madarame, H. 1, Seuberlich, T. 2, Vandevelde, M. 2, Zurbriggen, A. 2, and Oevermann, A. 2<br />

1. Veterinary Teaching Hospital, Azabu University, Japan<br />

2. Neurocenter, Vetsuisse Faculty Bern, Switzerland<br />

The interaction <strong>of</strong> internalin A, a major surface ligand <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

(LM), and its host cell-recep<strong>to</strong>r E-cadherin is required for the entry <strong>of</strong> LM in<strong>to</strong><br />

cells and for the crossing <strong>of</strong> the intestinal and placental barrier during infection. In<br />

ruminants, listeriosis occurs most commonly as rhombencephalitis, specifically<br />

targeting the brainstem, and it is believed that LM enters the brain via cranial<br />

nerves. However, the host cell recep<strong>to</strong>rs involved in brain invasion are not known.<br />

The aim <strong>of</strong> this study was <strong>to</strong> investigate the expression <strong>of</strong> E-cadherin in cattle<br />

brain, trigeminal nerve (TN) and its ganglion (TG), and thus its putative role in<br />

brain invasion. To this end, brains, TN and TG <strong>of</strong> cattle were examined by immunohis<strong>to</strong>chemistry<br />

(IHC, n = 6) and Western blot (n = 2). In the brain, strong<br />

membranous E-cadherin expression was observed in choroid plexus epithelial<br />

cells by IHC, whilst no other brain structures were labeled. TN and TG exhibited<br />

strong E-cadherin expression in Schwann cells and satellite cells (which are specialized<br />

Schwann cells), respectively. The labeling was particularly evident at the<br />

intercellular boundary <strong>of</strong> satellite cells and at the Schmidt-Lantermann incisures.<br />

Confirming the IHC results, E-cadherin specific bands <strong>of</strong> molecular masses <strong>of</strong> approximately<br />

37 kDa and 130 kD were evident in the WB <strong>of</strong> the TG, but were absent<br />

in the WB <strong>of</strong> the brainstem. These results indicate that E-cadherin expressing<br />

Schwann cells <strong>of</strong> cranial nerves might be the initial port <strong>of</strong> entry for LM in cattle<br />

rhombencephalitis. From t<strong>here</strong>, LM might gain access <strong>to</strong> the brain by cell-<strong>to</strong>-cell<br />

spread via axons. However, this hypothesis needs <strong>to</strong> be confirmed by complementary<br />

in vitro investigations <strong>of</strong> the molecular host-pathogen interactions.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Pathogenic potential <strong>of</strong> Listeria monocy<strong>to</strong>genes isolates<br />

from New Zealand seafood premises: implications for control<br />

Durante Cruz, C. and Fletcher, G.<br />

The New Zealand Institute for Plant and Food Research<br />

Listeria monocy<strong>to</strong>genes is likely <strong>to</strong> persist in food premises, leading <strong>to</strong> contamination<br />

<strong>of</strong> products processed t<strong>here</strong>in. It is t<strong>here</strong>fore important <strong>to</strong> assess the virulence potential<br />

<strong>of</strong> these isolates in order <strong>to</strong> evaluate the risk they can pose <strong>to</strong> consumers’<br />

heath. In this work, we assessed the invasion capacity <strong>of</strong> L. monocy<strong>to</strong>genes isolates<br />

obtained from New Zealand (NZ) seafood premises and correlated this with their<br />

persistence and genetic pr<strong>of</strong>ile. From 120 isolates obtained from four seafood premises,<br />

30 (28 from environmental and 2 from product samples) belonging <strong>to</strong> different<br />

serotypes and pulsotypes were compared in this study. These were compared with<br />

five human isolates obtained from the NZ Reference Culture Collection, two isolates<br />

(human and food) from a NZ seafood-related outbreak <strong>of</strong> listeriosis and one<br />

meat isolate shown <strong>to</strong> be resistant <strong>to</strong> high pressure conditions. ScottA was used as a<br />

reference strain and assigned an invasion index <strong>of</strong> 100 %. In vitro studies were conducted<br />

using Caco-2 cells (100:1 Listeria:Caco-2) followed by 40 min incubation at<br />

37 °C. Three independent experiments were performed on each isolate in duplicate.<br />

In Greenshell mussel premises, isolates from Plant I showed higher invasiveness<br />

than those from Plant II. All isolates were serotype 1/2a, except for one 1/2b from<br />

Plant I. The persistent and predominant pulsotype from Plant II had a low invasion<br />

index. Serotype 1/2b and 4b isolates from the two fish premises showed higher invasiveness<br />

than the serotype 1/2a isolates from mussel plants. Product (serotypes<br />

1/2a and 1/2b) and one human (serotype 1/2a) isolates showed higher invasion indices<br />

than environmental ones. The seafood outbreak strains had an intermediate<br />

invasion index. This study improves our understanding <strong>of</strong> L. monocy<strong>to</strong>genes present<br />

in NZ seafood premises and highlights the need <strong>to</strong> improve control. Although<br />

some environmental fac<strong>to</strong>ry isolates were potentially pathogenic, persistent strains<br />

were generally less invasive in Caco-2 cells.<br />

Copper homeostasis and virulence in Listeria monocy<strong>to</strong>genes<br />

David, C. 1, Schuler, S. 1, Glenn, S. 2, Jen, C. 1, Andrew, P. 2 and Roberts, I. S. 1<br />

1. University <strong>of</strong> Manchester, UK<br />

2. University <strong>of</strong> Leicester, UK<br />

Copper is essential for many bacteria being an important prosthetic group for certain<br />

enzymes. However at high levels copper is <strong>to</strong>xic in part due <strong>to</strong> its ability <strong>to</strong><br />

redox cycle and catalyse the formation <strong>of</strong> oxygen derived free radicals. As such<br />

bacteria need <strong>to</strong> be able <strong>to</strong> maintain copper homeostasis. In L. monocy<strong>to</strong>genes the<br />

problem <strong>of</strong> copper homeostasis is particularly acute. It inhabits a range <strong>of</strong> environments<br />

including soil, effluents and food w<strong>here</strong> it will be exposed <strong>to</strong> fluctuating<br />

copper levels. The use <strong>of</strong> copper as an anti-microbial in animal feeds, as a disinfectant<br />

in fac<strong>to</strong>ry-based farming and as a biocide in fruit production means L.<br />

monocy<strong>to</strong>genes will have <strong>to</strong> combat potentially <strong>to</strong>xic levels <strong>of</strong> copper during food<br />

production and processing. Adapting <strong>to</strong> fluctuations in the level <strong>of</strong> copper is important<br />

during infections. L. monocy<strong>to</strong>genes replicates in a number <strong>of</strong> host tissues<br />

that will <strong>of</strong>fer different challenges with regard <strong>to</strong> copper availability. Early in infection<br />

L. monocy<strong>to</strong>genes grows in the gall bladder, w<strong>here</strong> t<strong>here</strong> are high levels <strong>of</strong><br />

copper, w<strong>here</strong>as in the liver and spleen it replicates in an environment in which<br />

t<strong>here</strong> may be little free copper. Adapting <strong>to</strong> these different niches within the host<br />

requires effective copper homeostasis. In this paper we describe and analyse the<br />

operon that encodes for the only P1 type-ATPase copper exporter in L. monocy<strong>to</strong>genes<br />

and describe the molecular basis by which transcription <strong>of</strong> this operon is<br />

regulated by environmental copper and identify the role played by this operon in<br />

infection <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

R E F E R E N C E<br />

B / P<br />

53<br />

R E F E R E N C E<br />

B / P<br />

54


R E F E R E N C E<br />

B / P<br />

55<br />

R E F E R E N C E<br />

B / P<br />

56<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Pattern <strong>of</strong> cy<strong>to</strong>kine production during murine listeriosis<br />

Dussurget, O.<br />

Institut Pasteur, France<br />

Listeria monocy<strong>to</strong>genes causes listeriosis, an infection whose manifestations extend<br />

from gastroenteritis <strong>to</strong> septicemia, meningitis, encephalitis, abortions and<br />

perinatal infections. Cy<strong>to</strong>kines are important media<strong>to</strong>rs <strong>of</strong> host defense against<br />

pathogenic microorganisms. They control innate immune responses and contribute<br />

<strong>to</strong> shape adaptive immune responses. Several studies have previously analyzed<br />

the production <strong>of</strong> one or several cy<strong>to</strong>kines in response <strong>to</strong> Listeria monocy<strong>to</strong>genes<br />

in vitro and also in vivo. In this work, we used a murine model <strong>of</strong> primary<br />

listeriosis <strong>to</strong> systematically follow in the same animal the protein levels <strong>of</strong> 20 cy<strong>to</strong>kines<br />

in the blood and in the liver and spleen, two important organs in which L.<br />

monocy<strong>to</strong>genes replicate. Infection was shown <strong>to</strong> trigger the secretion <strong>of</strong> the major<br />

activa<strong>to</strong>r <strong>of</strong> macrophage antimicrobial activity, interferon (IFN) g, in blood, liver<br />

and spleen 24 h after intravenous inoculation. The early response <strong>to</strong> infection was<br />

also characterized by a significant increase in proinflamma<strong>to</strong>ry cy<strong>to</strong>kines tumor<br />

necrosis fac<strong>to</strong>r (TNF) a, interleukin (IL) 1a, IL1b and IL6, chemokines KC, IP10,<br />

MCP1, MIP1a and MIG, and growth fac<strong>to</strong>rs GM-CSF and V-EGF. Secretion <strong>of</strong> cy<strong>to</strong>kines<br />

important for the T cell response, e.g. IL2, and more specifically for T<br />

helper 1 (Th1) cell response, IL12 and IL10, was induced after infection. Surprisingly,<br />

the level <strong>of</strong> IL4 and IL13, two cy<strong>to</strong>kines involved in Th2 cell response, also<br />

increased upon infection. The level <strong>of</strong> most cy<strong>to</strong>kines increased in correlation with<br />

the number <strong>of</strong> bacteria in each organ during the first three days <strong>of</strong> listeriosis. In<br />

contrast, a strain <strong>of</strong> L. monocy<strong>to</strong>genes which does not produce the major virulence<br />

fac<strong>to</strong>r listeriolysin O (LLO), and is cleared very early from infected mice, failed <strong>to</strong><br />

stimulate cy<strong>to</strong>kine production. Together, this first multiplex analysis <strong>of</strong> cy<strong>to</strong>kine<br />

production in vivo represents an important basis <strong>to</strong> identify virulence determinants<br />

involved in the modulation <strong>of</strong> host immune response.<br />

Analysis <strong>of</strong> the post-translocation chaperone PrsA2 and its unique role<br />

in facilitating Listeria monocy<strong>to</strong>genes pathogenesis<br />

Alonzo, F. and Freitag, N.<br />

University Of Illinois at Chicago, USA<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive bacterial pathogen whose ability <strong>to</strong><br />

cause disease is dependent upon the coordinated expression and activity <strong>of</strong> secreted<br />

products that promote virulence. We have shown that the post-translocation<br />

chaperone PrsA2 is critical for L. monocy<strong>to</strong>genes pathogenesis and that prsA2<br />

mutants exhibit a substantial cell-<strong>to</strong>-cell spread defect in tissue culture. We t<strong>here</strong>fore<br />

hypothesized that a loss <strong>of</strong> PrsA2 results in perturbation and/or altered activity<br />

<strong>of</strong> fac<strong>to</strong>rs responsible for intracellular survival. PrsA2 was found <strong>to</strong> promote<br />

the activity and stability <strong>of</strong> two secreted virulence fac<strong>to</strong>rs: listeriolysin O (LLO)<br />

and the broad range phospholipase PC-PLC. Culture supernatants from strains<br />

containing deletions <strong>of</strong> prsA2 exhibited reduced hemolytic activity as measured<br />

by the lysis <strong>of</strong> red blood cells in vitro, and this reduced activity appeared <strong>to</strong> be the<br />

result <strong>of</strong> altered stability and increased proteolytic cleavage <strong>of</strong> LLO. prsA2 mutants<br />

were also defective for processing <strong>of</strong> PC-PLC from its pro-form <strong>to</strong> its enzymatically<br />

active mature form, resulting in decreased phospholipase activity in vitro<br />

as well as within infected host cells. PrsA2’s role in bacterial virulence was found <strong>to</strong><br />

be unique from the related L. monocy<strong>to</strong>genes chaperone PrsA1, a protein that<br />

shares a high degree <strong>of</strong> sequence similarity with PrsA2 but makes no discernable<br />

contribution <strong>to</strong> pathogenesis. Proteomic analysis <strong>of</strong> L. monocy<strong>to</strong>genes secreted<br />

proteins suggests that PrsA2 is required for the correct localization <strong>of</strong> additional<br />

bacterial gene products. Taken <strong>to</strong>gether, these data indicate that PrsA2 has been<br />

functionally adapted <strong>to</strong> promote the secretion and activity <strong>of</strong> multiple L. monocy<strong>to</strong>genes<br />

fac<strong>to</strong>rs that play important roles in bacterial virulence.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

CtaP is a multifunctional cysteine-transport associated protein<br />

required for Listeria monocy<strong>to</strong>genes pathogenesis<br />

Xayarath, B.<br />

University <strong>of</strong> Illinois at Chicago, USA<br />

The bacterial pathogen Listeria monocy<strong>to</strong>genes survives under a myriad <strong>of</strong> conditions<br />

in the outside environment and also within the human host w<strong>here</strong> infections<br />

can result in severe disease. Bacterial life within the host requires the expression<br />

<strong>of</strong> genes with roles in nutrient acquisition as well as the biosynthesis <strong>of</strong> bacterial<br />

products required <strong>to</strong> support intracellular growth. A gene product identified as<br />

the substrate-binding component <strong>of</strong> a novel oligopeptide transport system (encoded<br />

by lmo0135) was recently shown <strong>to</strong> be required for L. monocy<strong>to</strong>genes virulence.<br />

We report <strong>here</strong> that the gene product encoded by lmo0135 makes multiple<br />

contributions <strong>to</strong> L. monocy<strong>to</strong>genes growth and survival in environments both inside<br />

and outside <strong>of</strong> host cells. The lmo0135 gene product was required for bacterial<br />

growth in the presence <strong>of</strong> low concentrations <strong>of</strong> cysteine in vitro, suggesting<br />

lmo0135 and its associated transport system function in high affinity cysteine<br />

transport. We have t<strong>here</strong>fore designated the lmo0135 gene products as CtaP for<br />

cysteine-transport associated protein. Interestingly, CtaP was not required for<br />

bacterial replication within the host cy<strong>to</strong>sol, but was required for bacterial adhesion<br />

implicating a role for this protein in bacterial attachment <strong>to</strong> host cells. CtaP<br />

appears <strong>to</strong> function directly as an adhesin as latex beads coupled <strong>to</strong> CtaP were<br />

found <strong>to</strong> specifically bind epithelial cell monolayers in tissue culture assays. In<br />

addition, loss <strong>of</strong> CtaP increased membrane permeability and acid sensitivity, resulted<br />

in altered bacterial surface hydrophobicity, and severely attenuated virulence<br />

following both intragastric and intravenous inoculation <strong>of</strong> mice. Taken <strong>to</strong>gether,<br />

the data presented indicates that CtaP is part <strong>of</strong> a unique ABC transport<br />

system that contributes <strong>to</strong> multiple facets <strong>of</strong> L. monocy<strong>to</strong>genes physiology, growth,<br />

and survival both inside and outside <strong>of</strong> animal cells.<br />

Enzymatic activity <strong>of</strong> the metalloprotease <strong>of</strong> Listeria is regulated by pH<br />

Forster, B. M.*, Pavinski Bitar, A., Slepkov, E. R. and Marquis, H.<br />

Cornell University, USA<br />

The broad-range phospholipase C <strong>of</strong> Listeria monocy<strong>to</strong>genes, PC-PLC, contributes<br />

<strong>to</strong> escape from vacuoles. During infection, PC-PLC accumulates at the membrane<br />

cell wall interface until a decrease in vacuolar pH triggers its proteolytic maturation<br />

and release. PC-PLC maturation is mediated by the metalloprotease <strong>of</strong> Listeria<br />

(Mpl), which is produced as a zymogen and matures via intramolecular au<strong>to</strong>catalysis.<br />

We tested the hypothesis that Mpl activity is regulated by pH. Radiolabeled<br />

infected cells were perfused <strong>to</strong> manipulate host cell cy<strong>to</strong>solic pH, and secreted Mpl<br />

was immunoprecipitated. Mature Mpl was detected in samples treated at pH 6.5,<br />

but not at pH 7.3. To distinguish whether pH regulates au<strong>to</strong>catalysis or cell wall<br />

translocation <strong>of</strong> bacterium-associated Mpl, we inserted a Flag tag at the N-terminus<br />

<strong>of</strong> the pro or catalytic domain <strong>of</strong> Mpl, and used an antibody that recognizes only<br />

N-terminal Flag <strong>to</strong> detect Mpl. Results from fluorescence microscopy indicated<br />

that the zymogen is bacterium-associated at physiological pH, but not following<br />

acidification <strong>of</strong> the host cell cy<strong>to</strong>sol. Mature Mpl was not detected bacterium-associated<br />

at either pH. Lastly, using purified mature Mpl and the pr<strong>of</strong>orm <strong>of</strong> PC-<br />

PLC, we determined that processing <strong>of</strong> the PC-PLC propeptide by mature Mpl occurs<br />

only at acidic pH. Together, these results indicated that Mpl enzymatic activity<br />

is regulated by pH, whether it undergoes au<strong>to</strong>catalysis or mediates the maturation<br />

<strong>of</strong> PC-PLC. Future studies will aim at assessing the possibility that specific amino<br />

acid residues act as pH sensors in the regulation <strong>of</strong> Mpl au<strong>to</strong>catalysis.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

B / P<br />

57<br />

R E F E R E N C E<br />

B / P<br />

58


R E F E R E N C E<br />

B / P<br />

59<br />

R E F E R E N C E<br />

B / P<br />

60<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Identification <strong>of</strong> propeptide residues regulating the<br />

compartmentalization, maturation, and activity <strong>of</strong> the broad-range<br />

phospholipase C <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Slepkov, E. R., Pavinski Bitar, A. and Marquis, H.<br />

Cornell University, USA<br />

The broad-range phospholipase C <strong>of</strong> Listeria monocy<strong>to</strong>genes, PC-PLC, is made as<br />

an inactive proenzyme whose activation is regulated by pH and by the metalloprotease<br />

<strong>of</strong> Listeria (Mpl). The propeptide <strong>of</strong> PC-PLC is comprised <strong>of</strong> 24 amino<br />

acid residues (C28-S51) and regulates PC-PLC compartmentalization and activity.<br />

In this study, we generated a series <strong>of</strong> propeptide mutants <strong>to</strong> determine the minimal<br />

requirement <strong>to</strong> prevent PC-PLC activity, and <strong>to</strong> identify amino acid residues<br />

regulating compartmentalization and maturation. To determine how many<br />

residues were required <strong>to</strong> prevent PC-PLC activity, nested deletions were generated<br />

in the PC-PLC propeptide <strong>of</strong> an Mpl-minus strain and phospholipase activity<br />

was assessed on egg yolk plates. We found that a single amino acid residue at position<br />

P1 (S51) <strong>of</strong> the cleavage site is sufficient <strong>to</strong> prevent activity, which is consistent<br />

with P1’ (W52) being located within the active site pocket. The requirements <strong>to</strong><br />

retain the pr<strong>of</strong>orm <strong>of</strong> PC-PLC bacterium-associated were assessed by immunoprecipitating<br />

secreted PC-PLC from radiolabeled infected cells maintained at<br />

physiological pH, w<strong>here</strong>as the influence <strong>of</strong> the propeptide on the ability <strong>of</strong> Mpl <strong>to</strong><br />

mediate PC-PLC maturation was assessed by immunoprecipitating secreted PC-<br />

PLC from radiolabeled infected cells perfused with buffer at acidic pH. We observed<br />

that the triple mutant E31A Y32A L33A is translocated across the cell wall<br />

more effectively than wild-type at physiological pH, and that Y32A and L33A single<br />

point mutants are less effectively processed by Mpl at acidic pH. These results<br />

indicated that the N-terminus <strong>of</strong> the propeptide regulates PC-PLC compartmentalization<br />

and the ability <strong>of</strong> Mpl <strong>to</strong> process PC-PLC. Considering that Y32 and L33<br />

are located 19 and 18 residues upstream <strong>of</strong> the propeptide cleavage site, we suggest<br />

that these two residues initiate the interaction with mature Mpl leading <strong>to</strong> processing<br />

<strong>of</strong> the propeptide at S51.<br />

A mouse model <strong>of</strong> fe<strong>to</strong>placental Listeria monocy<strong>to</strong>genes infection<br />

and abortion<br />

Poulsen, K. P., Faith, N., Laura Knoll, L. and Czuprynski, C.<br />

School <strong>of</strong> Veterinary Medicine, University <strong>of</strong> Wisconsin Madison, USA<br />

Foodborne outbreaks <strong>of</strong> Listeria monocy<strong>to</strong>genes, particularly with serotype 4b,<br />

continue<strong>to</strong> occur. Pregnant women are over represented in listeriosis outbreaks<br />

(17-fold increasein incidence <strong>of</strong> disease). Infection <strong>of</strong> the fetus and placenta results<br />

in abortion, stillbirth,or premature parturition. The latter carries with it a<br />

high risk for neonatal sepsis andmeningitis. Alterations in cellular and molecular<br />

components <strong>of</strong> the immune systemoccur during pregnancy <strong>to</strong> prevent rejection<br />

<strong>of</strong> the maternal allograft (i.e. fetus). How these pregnancy associated changes affect<br />

vertical transmission <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> the fetus, is poorly unders<strong>to</strong>od.<br />

We have developed a mouse model for infection with aserotype 4b strain <strong>of</strong> L.<br />

monocy<strong>to</strong>genes isolated from a foodborne listeriosis outbreak thatresulted in abortion<br />

and fetal death. Intragastric infection <strong>of</strong> pregnant mice resulted insevere infection<br />

<strong>of</strong> the maternal and fetal tissues, in both C57BL/6J and A/J mousestrains.<br />

Use <strong>of</strong> a luciferase tagged L. monocy<strong>to</strong>genes strain and bioluminescent technology<br />

allowed us <strong>to</strong> visualize infection <strong>of</strong> maternal and fetal tissues and shedding <strong>of</strong> L.<br />

monocy<strong>to</strong>genes cells in feces and vaginal secretions <strong>of</strong> aborting mice. These data<br />

support the use <strong>of</strong> pregnant mice as a model for maternal and fetal L. monocy<strong>to</strong>genes<br />

infection. This model will prove useful in future investigations <strong>of</strong> the maternal<br />

and fetal immune response <strong>to</strong> listeriosis during pregnancy.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Listeria infection <strong>of</strong> the insect model system Galleria mellonella<br />

Joyce, S. A. and Gahan, C. G.<br />

APC and department <strong>of</strong> Microbiology UCC, Ireland<br />

With the exception <strong>of</strong> mice, alternative model systems are constrained by an upper<br />

temperature limit for their survival. This temperature varies from 25 °C for Acanthamoeba<br />

and Danio rerio <strong>to</strong> 28 °C for C. elegans and for Drosophila species. This<br />

limitation questions the relevance <strong>of</strong> these models for use with human pathogens<br />

w<strong>here</strong> virulence gene expression is temperature dependent. Galleria mellonella,<br />

the Greater Wax Moth Larvae, is well documented as an infection model amenable<br />

<strong>to</strong> elevated temperature <strong>of</strong> 37 °C. Here, we examine and establish Galleria mellonella<br />

as a relevant model for infection by Listeria species. Mutants attenuated for<br />

virulence in mice are similarly attenuated in this model system. We demonstrate<br />

that haemolysin production is the dominant fac<strong>to</strong>r in the pathogenic process <strong>of</strong> L.<br />

monocy<strong>to</strong>genes <strong>to</strong> G. mellonella. Real time visualization <strong>of</strong> infection reveals that the<br />

same sub-sets <strong>of</strong> virulence genes required for L. monocy<strong>to</strong>genes infection <strong>of</strong> both<br />

humans and mice are expressed (differentially) and required for insect infections.<br />

L. monocy<strong>to</strong>genes replication occurs in G. mellonella. On examining the host response<br />

we find that L. monocy<strong>to</strong>genes is mainly internalized within one hour <strong>of</strong> infection.<br />

In the presence <strong>of</strong> hemolysin, hemocyte viability decreased eight- fold and<br />

in the absence <strong>of</strong> PrfA a four-fold increase in viability was evident relative <strong>to</strong> WT<br />

strain. Bacterial load in hemocytes was significantly increased on infection with a<br />

mutant in hemolysin (DhlyA) production and a mutant in stress fac<strong>to</strong>r production<br />

(DsigB) compared <strong>to</strong> wt infection. We found that the Phenyloxidase (PO) system<br />

mounts a response <strong>to</strong> the presence <strong>of</strong> Listeria species within 4 hour <strong>of</strong> insect infection<br />

and that infection by L. monocy<strong>to</strong>genes results in the production <strong>of</strong> anti microbial<br />

peptides (AMPs). Taken <strong>to</strong>gether, we propose that G. mellonella is a relevant<br />

model for infection by L. monocy<strong>to</strong>genes at 37 °C.<br />

Construction <strong>of</strong> a murinised Listeria monocy<strong>to</strong>genes H7858 (4b) strain<br />

for improved murine infection<br />

Cummins, J. and Gahan, C.<br />

Science Foundation Ireland<br />

Listeria monocy<strong>to</strong>genes is Gram-positive food borne pathogen that is capable <strong>of</strong><br />

causing listerosis culminating in various diseases in humans such as gastroenteritis,<br />

meningitis, encephalitis and spontaneous abortions. L. monocy<strong>to</strong>genes is capable<br />

<strong>of</strong> internalisation in<strong>to</strong> non-phagocytic cells and crossing the intestinal, the<br />

placental and the blood-brain barriers. Internalisation in<strong>to</strong> non-pr<strong>of</strong>essional<br />

phagocytic cells is mediated in particular by the surface protein, InlA. InlA promotes<br />

listerial uptake in<strong>to</strong> enterocytes by targeting the N-terminal domain <strong>of</strong> the<br />

human E-cadherin. However, this interaction is species specific with its interaction<br />

completely impaired in the rat and mouse models due <strong>to</strong> the absence <strong>of</strong> a proline<br />

at position 16 <strong>of</strong> the N-terminal E-cadherin repeat. A landmark result was the<br />

development <strong>of</strong> a transgenic mouse line expressing human E-cadherin by intestinal<br />

enterocytes. However, an alternative approach <strong>to</strong> overcome the lack <strong>of</strong> appropriate<br />

animal models has been the creation <strong>of</strong> a murinised strain <strong>of</strong> L. monocy<strong>to</strong>genes<br />

EGDe (Wollert et al., 2007 Cell. 129(5):891-902). By the substitution <strong>of</strong><br />

two amino acids within the InlA protein we have created a murnised strain in the<br />

H7858 4b background. Our results have demonstrated that this mutant has an increased<br />

ability <strong>to</strong> infect mice by the oral route and does not increase invasion<br />

within human cells. The benefits <strong>of</strong> this approach will be discussed.<br />

R E F E R E N C E<br />

B / P<br />

61<br />

R E F E R E N C E<br />

B / P<br />

62


R E F E R E N C E<br />

B / P<br />

63<br />

R E F E R E N C E<br />

B / P<br />

64<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

The role <strong>of</strong> a phosphoinositide phosphatase in the intracellular survival<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Wang, J., Corbett, D. and Roberts, I. S.<br />

Faculty <strong>of</strong> Life Sciences, University <strong>of</strong> Manchester, UK<br />

L. monocy<strong>to</strong>genes is capable <strong>of</strong> invading and growing in a number <strong>of</strong> host cells. Following<br />

uptake it escapes from the phagosome and grows in the cy<strong>to</strong>sol recruiting<br />

actin filaments <strong>to</strong> move and spread from cell <strong>to</strong> cell. Once inside the cell it subverts<br />

the host cell physiology <strong>to</strong> promote its own growth and survival. Phosphoinositides<br />

are phospholipids present in host cell membranes that play key regula<strong>to</strong>ry<br />

roles in orchestrating cell physiology. They control a broad range <strong>of</strong> processes<br />

including organelle identity, signal transduction and cell migration. The basis <strong>of</strong><br />

this control is their relative state <strong>of</strong> phosphorylation that is controlled through<br />

the action <strong>of</strong> host kinases and phosphatases. The central role played by phosphoinositides<br />

inside cells make them a prime target for subversion by intracellular<br />

pathogens wishing <strong>to</strong> hijack the cell’s physiology. Recently we discovered that L.<br />

monocy<strong>to</strong>genes expresses a phosphoinositide phosphatase, capable <strong>of</strong> removing<br />

phosphate groups from a number <strong>of</strong> important phosphoinositides. Importantly<br />

we demonstrated that this enzyme was essential for the growth <strong>of</strong> L. monocy<strong>to</strong>genes<br />

inside infected cells and probably is important in escaping the phagosome.<br />

This is the first identification <strong>of</strong> a phosphoinositide phosphatase enzyme playing<br />

a key role in the intracellular survival <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

Virulence gene expression in Listeria monocy<strong>to</strong>genes strains isolated<br />

from different sources<br />

Alessandria, V.<br />

University <strong>of</strong> Turin, Italy<br />

Listeria monocy<strong>to</strong>genes is an important food-borne pathogen that has the capacity<br />

<strong>to</strong> cause severe infections. Ubiqui<strong>to</strong>us microorganism, it is commonly isolated<br />

from foods <strong>of</strong> animal origin, mainly meat and milk products. However, due <strong>to</strong> its<br />

capacity <strong>to</strong> develop at refrigeration temperatures, human listeriosis outbreaks are<br />

most <strong>of</strong>ten associated with ready-<strong>to</strong>-eat food products that are consumed without<br />

prior cooking. The incidence <strong>of</strong> the disease depends on different fac<strong>to</strong>rs including<br />

the infective dose and the immunity conditions <strong>of</strong> the host. The present<br />

work focuses on the expression analysis <strong>of</strong> four virulence genes (sigB, plcA, hly<br />

and iap) in 11 different strains <strong>of</strong> L. monocy<strong>to</strong>genes and, more in detail, 3 collection<br />

strains (EGDe, NCTC10527, SCOTT A), 7 isolated from food matrices (4 <strong>of</strong> which<br />

isolated from meat products and three from diary products) and 1 isolated from<br />

humans. In the first step, the expression analysis was performed in vitro, culturing<br />

each strain in BHI (Brain Heart Infusion) medium, in order <strong>to</strong> identify possible<br />

differences in the expression level for the different strains. The combined use <strong>of</strong><br />

reverse transcription and quantitative PCR (qRT-PCR) was used <strong>to</strong> evaluate gene<br />

expression. As a second step, we wanted <strong>to</strong> evaluate the trend <strong>of</strong> expression <strong>of</strong> the<br />

genes in food. Analyses were performed inoculating L. monocy<strong>to</strong>genes in milk at<br />

different conditions and in particular at two temperatures (4 and 12 °C) for different<br />

times (24 and 48 hours). Significant expression differences emerged for the<br />

different genes, without showing any significant association between the expression<br />

<strong>of</strong> the genes and the origin <strong>of</strong> the different strains.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Targeted Signature-tagged mutagenesis for phenotype screening<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes mutants<br />

Henriques, A., Carvalho, F. and Cabanes, D.<br />

IBMC, Portugal<br />

Listeria monocy<strong>to</strong>genes is a foodborne facultative intracellular human pathogen<br />

possessing a wide range <strong>of</strong> virulence fac<strong>to</strong>rs tailored for pathogenesis. Common<br />

strategies used <strong>to</strong> identify new virulence fac<strong>to</strong>rs are based on mutagenesis and in<br />

vivo phenotype analysis, which requires testing <strong>of</strong> generated mutants versus wild<br />

type strain in separate mice cohorts. To streamline this analysis, a targeted signature-tagged<br />

mutagenesis (T-STM) method was developed. Based on the concept <strong>of</strong><br />

STM, w<strong>here</strong> random mutants identifiable by specific tagging are negatively selected,<br />

we developed a method that enables specific gene targeting in L. monocy<strong>to</strong>genes<br />

genome for the creation <strong>of</strong> insertional mutants and their simultaneous<br />

phenotypic analysis in mice. Tagged “wild type” and mutants for prfA and two unknown<br />

LPXTG genes were constructed by inserting differentially tagged suicide<br />

vec<strong>to</strong>rs. T-STM uses defined tags that allow the discrimination between taggedstrains<br />

in a pool by PCR. The four tagged-strains were pooled in equivalent<br />

amounts in<strong>to</strong> one single inoculum (Input) used for oral and intravenous infection<br />

<strong>of</strong> mice. Three days after infection, intestine, spleen, and liver <strong>of</strong> mice were collected,<br />

homogenised and plated on rich medium. Resultant bacterial growth (Output)<br />

was used as template for PCR screening and evaluation <strong>of</strong> the relative proportion<br />

<strong>of</strong> each tagged-strain. Mutants were considered less virulent when<br />

undetectable (or less detectable) by PCR in the output. This technique allowed us<br />

<strong>to</strong> clearly distinguish the severely attenuated prfA mutant, validating the method.<br />

PCR products for the two LPXTG protein-encoding mutants appeared in equal<br />

proportion <strong>to</strong> the wild type strain in the output, suggesting that these proteins are<br />

not involved in Listeria virulence in this infectious model. This study shows the<br />

feasibility and advantage <strong>of</strong> the developed T-STM method for simultaneous virulence<br />

analysis <strong>of</strong> pools <strong>of</strong> Listeria tagged mutants. This approach should be now<br />

used for larger-scale phenotypic analysis <strong>of</strong> specific L. monocy<strong>to</strong>genes mutants.<br />

Listeria monocy<strong>to</strong>genes cellular infection triggers tyrosinephosphorylation<br />

<strong>of</strong> Myosin IIA, a new protein involved in invasion<br />

Almeida, M. T., Cabanes, D. and Sousa, S.<br />

IBMC, Portugal<br />

Listeria monocy<strong>to</strong>genes is a human food borne pathogen that may lead, in particular<br />

in immunocompromised individuals, <strong>to</strong> a severe disease characterized by septicemias,<br />

meningitis, meningo-encephalitis and abortions. The study <strong>of</strong> the cell<br />

biology <strong>of</strong> the Listeria infectious process provided insights in the way bacteria manipulate<br />

the host and revealed unsuspected functions <strong>of</strong> cellular proteins. To cause<br />

infection pathogens interfere with crucial host intracellular pathways, and different<br />

pathogens <strong>of</strong>ten hijack the same signaling pathways. In particular, host phosphorylation<br />

cascades are preferential targets <strong>of</strong> infecting bacteria. In this study,<br />

using L. monocy<strong>to</strong>genes as a pathogen model, we showed that eukaryotic cells present<br />

a variable protein phosphorylation pattern upon infection. We addressed in<br />

particular the tyrosine-phosphorylated protein pr<strong>of</strong>ile triggered by Listeria infection<br />

and identified the mo<strong>to</strong>r protein, Myosin IIA (MyoIIA), as differentially tyrosine-phosphorylated<br />

in response <strong>to</strong> Listeria uptake. We demonstrated that MyoIIA<br />

is not only tyrosine-phosphorylated over the time <strong>of</strong> infection, but is also<br />

recruited with actin at the bacteria entry site. In addition, we were able <strong>to</strong> show<br />

that the inhibition <strong>of</strong> MyoIIA activity affected Listeria entry in<strong>to</strong> non-phagocytic<br />

cells. The reduction <strong>of</strong> MyoIIA expression using RNAi techniques resulted in an<br />

increased Listeria uptake. Together these data point <strong>to</strong> the role <strong>of</strong> a novel myosin<br />

class in the internalization <strong>of</strong> Listeria, correlating for the first time, myosin posttranslational<br />

modifications and Listeria infection.<br />

R E F E R E N C E<br />

B / P<br />

65<br />

R E F E R E N C E<br />

B / P<br />

66


R E F E R E N C E<br />

B / P<br />

67<br />

R E F E R E N C E<br />

B / P<br />

68<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Invasion pr<strong>of</strong>ile <strong>of</strong> Listeria monocy<strong>to</strong>genes strains involved<br />

in invasive and gastroenteritis listeriosis outbreaks<br />

Laksanalamai, P., Sahu, S. and Datta, A.<br />

Food and Drug Administration, USA<br />

Listeria monocy<strong>to</strong>genes (Lm), the causative agent <strong>of</strong> foodborne human listeriosis, is<br />

a serious public health concern. The disease is characterized by meningitis, septicemia,<br />

abortion and death in immuno-compromised population. Growing numbers<br />

<strong>of</strong> evidence have revealed that Lm can also cause self-limited febrile gastroenteritis<br />

in healthy individuals. To understand the mechanisms underlying two<br />

different disease outcomes several efforts have been made <strong>to</strong> differentiate gastroenteritis<br />

and invasive strains <strong>of</strong> Lm. Since the difference between these two<br />

types <strong>of</strong> outbreaks is Listerial ability <strong>to</strong> invade, in this study we have examined<br />

the expression <strong>of</strong> two internalin genes, inlA and inlB which encode virulence proteins<br />

required for the internalization process. The inlA and inlB gene expression<br />

pr<strong>of</strong>iles were determined by quantitative RT-PCR from several strains representing<br />

two groups <strong>of</strong> Lm that cause either invasive listeriosis or febrile gastroenteritis.<br />

Analysis <strong>of</strong> the inlA and inlB gene expression from food and patient isolates<br />

showed similar levels <strong>of</strong> expression from isolates originated from the same outbreaks.<br />

Surprisingly, the inlA and inlB gene expression from Lm isolated from invasive<br />

listeriosis outbreaks appears <strong>to</strong> have lower expression that those that<br />

caused gastroenteritis listeriosis. In contrast the expression <strong>of</strong> iap, a gene involved<br />

in host cell invasion, was higher in invasive strains. To understand the correlation<br />

between the inlA and inlB gene expression and invasion <strong>of</strong> Lm, we also investigated<br />

the efficiency <strong>of</strong> invasion <strong>of</strong> these strains by in vitro invasion assay using eukaryotic<br />

cell lines, Caco2 and HepG2. The results appeared <strong>to</strong> correlate with the<br />

inlA and inlB expression in that the invasion efficiency is higher in the gastroenteritis<br />

strains. Global gene expression pr<strong>of</strong>iles <strong>of</strong> these strains are currently being<br />

investigated using a DNA microarray.<br />

Molecular characterization <strong>of</strong> the Vip-Gp96 interaction<br />

Martins, M., Cabanes, D. and Sousa, S.<br />

IBMC, Portugal<br />

The study <strong>of</strong> mechanisms exploited by Listeria monocy<strong>to</strong>genes <strong>to</strong> cause infection<br />

provided new insights in the way bacteria manipulate the host cell machinery. To<br />

subvert the host cell signalling pathways, Listeria uses a complex set <strong>of</strong> surface<br />

proteins called virulence fac<strong>to</strong>rs that act in concert <strong>to</strong> allow bacterial infection<br />

and evasion from the host immune system. Recently, we reported that Vip, a L.<br />

monocy<strong>to</strong>genes virulence fac<strong>to</strong>r interacts with Gp96, an endoplasmic reticulum<br />

(ER) resident chaperone that in cell stress conditions is targeted <strong>to</strong> cell surface,<br />

promoting host cell invasion. The purpose <strong>of</strong> this study was the identification <strong>of</strong><br />

the Gp96 domain involved in the interaction with Vip. Using protein-protein interaction<br />

and invasion assays, we demonstrated that the region 22-411 amino acids<br />

<strong>of</strong> Gp96 (N-terminal) seems <strong>to</strong> be exposed outside <strong>of</strong> the cell thus available for<br />

Vip binding. This interaction is required for Listeria uptake in<strong>to</strong> host cells. Although<br />

this entire region participates in the interaction, the domain comprising<br />

the 22-192 amino acids seemed <strong>to</strong> be the one crucial for the interaction. The existence<br />

<strong>of</strong> cell signalling events downstream Vip-Gp96 interaction was also investigated.<br />

Signal transducer and activa<strong>to</strong>r <strong>of</strong> transcription 3 (Stat3) activation during<br />

L. monocy<strong>to</strong>genes invasion was addressed, suggesting that Stat3 was not activated<br />

in response <strong>to</strong> Listeria entry. In addition, we demonstrated that Vip seems <strong>to</strong> have<br />

no role in Listeria phagocy<strong>to</strong>sis by bone marrow-derived macrophages (BMMØ)<br />

and in bacterial survival within them.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Investigation <strong>of</strong> the molecular mechanisms by which<br />

Listeria monocy<strong>to</strong>genes grows in the mammalian gall bladder<br />

Dowd, G., Joyce, S., Casey, P. G., Hill, C. and Gahan, C. G.<br />

University College Cork, Ireland<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen that is a significant cause <strong>of</strong> mortalities<br />

due <strong>to</strong> consumption <strong>of</strong> contaminated foods. Recent work has demonstrated<br />

that the pathogen grows in the mammalian gall bladder and that gall bladder growth<br />

forms a significant phase <strong>of</strong> the infectious process. However, relatively little is<br />

known about the mechanisms that underpin the growth <strong>of</strong> the pathogen in this milieu.<br />

Here we have performed basic experiments which demonstrate that L. monocy<strong>to</strong>genes<br />

grows well in bile from the porcine gall bladder and can utilize bile as a<br />

source <strong>of</strong> energy and nutrients. We used bioluminescence-labeled L. monocy<strong>to</strong>genes<br />

<strong>to</strong> allow visualization <strong>of</strong> bacterial growth kinetics in porcine gall bladders infected<br />

ex vivo. We then performed random mutagenesis <strong>of</strong> L. monocy<strong>to</strong>genes and selected<br />

mutants that grow well under labora<strong>to</strong>ry conditions but fail <strong>to</strong> grow in bile. Sequencing<br />

<strong>of</strong> the genes mutated in this experiment demonstrated that the pathogen<br />

requires genetic loci that are involved in central metabolism for growth in bile, including<br />

genes encoding proteins required for amino acid biosynthesis and biotin<br />

metabolism. The majority <strong>of</strong> mutants that were unable <strong>to</strong> grow in bile were also<br />

attenuated for virulence in the mouse model <strong>of</strong> infection. Overall the work is the<br />

first <strong>to</strong> demonstrate the requirement for specific gene sets for the growth <strong>of</strong> L.<br />

monocy<strong>to</strong>genes in the specialized environment <strong>of</strong> the mammalian gall bladder.<br />

Role <strong>of</strong> cadmium efflux system in Listeria monocy<strong>to</strong>genes virulence<br />

Camejo, A. and Cabanes, D.<br />

IBMC, Portugal<br />

Listeria monocy<strong>to</strong>genes is a human intracellular pathogen able <strong>to</strong> colonize host<br />

tissues after ingestion <strong>of</strong> contaminated food, causing severe invasive infections.<br />

In order <strong>to</strong> gain a better understanding <strong>of</strong> the nature <strong>of</strong> host–pathogen interactions,<br />

we studied the L. monocy<strong>to</strong>genes genome expression during mouse infection.<br />

We found that the shift <strong>of</strong> the Listeria genome expression during infection is<br />

characterized by the activation <strong>of</strong> genes involved in virulence, stress, subversion <strong>of</strong><br />

the host immune system, and by the adaptation <strong>of</strong> the bacterial metabolism <strong>to</strong><br />

host conditions. Mutagenesis <strong>of</strong> genes highly induced in vivo allowed the identification<br />

<strong>of</strong> novel L. monocy<strong>to</strong>genes virulence fac<strong>to</strong>rs, including cadC. Cadmium resistance<br />

in Listeria and other gram positive bacteria is an energy-dependent cadmium<br />

efflux system, involving two proteins, CadA and CadC. CadA has been shown<br />

<strong>to</strong> be a cadmium efflux P-type ATPase; cadC encodes a transcriptional regula<strong>to</strong>r<br />

that is a member <strong>of</strong> the ArsR metalloregula<strong>to</strong>ry proteins. The strong in vivo activation<br />

<strong>of</strong> cadC and the significant impaired virulence <strong>of</strong> the cadC mutant suggest<br />

that this heavy metal resistance system constitutes an advantage for in vivo Listeria<br />

survival. In addition, this system has never been previously implicated in the<br />

virulence <strong>of</strong> other pathogens. The detailed regulation <strong>of</strong> the CadAC system role as<br />

well as its role in cadmium resistance and in the Listeria infectious process will be<br />

presented and discussed.<br />

R E F E R E N C E<br />

B / P<br />

69<br />

R E F E R E N C E<br />

B / P<br />

70


R E F E R E N C E<br />

B / P<br />

71<br />

R E F E R E N C E<br />

B / P<br />

72<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Investigation <strong>of</strong> chitinase as a potential virulence fac<strong>to</strong>r<br />

in Listeria monocy<strong>to</strong>genes<br />

Chaudhuri, S.<br />

University <strong>of</strong> Illinois in Chicago, USA<br />

Listeria monocy<strong>to</strong>genes is an environmental pathogen that causes disease in humans<br />

primarily following the consumption <strong>of</strong> contaminated food products. While<br />

considerable attention has been given <strong>to</strong> the identification <strong>of</strong> bacterial virulence<br />

fac<strong>to</strong>rs specifically expressed within host cells, relatively little is known regarding<br />

gene products that may contribute <strong>to</strong> bacterial life in the outside environment<br />

while also potentially serving a role for virulence within the host. L. monocy<strong>to</strong>genes<br />

has been shown <strong>to</strong> produce and secrete two chitinases, ChiA and ChiB, as<br />

well as a chitin binding protein encoded by lmo2467. ChiA and ChiB are chitin degrading<br />

enzymes that are presumed <strong>to</strong> facilitate L. monocy<strong>to</strong>genes life in the outside<br />

environment w<strong>here</strong> chitin is plentiful. Recently however, chitinases and chitin<br />

binding proteins have been reported <strong>to</strong> contribute <strong>to</strong> the pathogenicity <strong>of</strong> two unrelated<br />

environmental bacterial pathogens, Legionella pneumophila and Vibrio<br />

cholerae. The chitinases produced by these bacteria appear <strong>to</strong> enhance bacterial<br />

colonization <strong>of</strong> mouse lung (L. pneumophila) and intestine (V. cholerae). To determine<br />

if chitinases or chitin binding proteins contribute <strong>to</strong> L. monocy<strong>to</strong>genes pathogenecity,<br />

in-frame deletion mutants were constructed in chiA, chiB, and lmo2467,<br />

and a triple gene deletion mutant was also constructed. All mutants were found <strong>to</strong><br />

grow similarly <strong>to</strong> wild type L. monocy<strong>to</strong>genes in BHI broth culture and within infected<br />

tissue culture cells. However, following intravenous infection <strong>of</strong> mice, two <strong>of</strong><br />

the single mutants, ∆chiA and ∆chiB, as well as the triple mutant, ∆chiA ∆chiB<br />

∆lmo2467 were found <strong>to</strong> be defective for bacterial growth in the livers and spleens<br />

<strong>of</strong> infected animals. As mammals do not produce chitin, these results suggest that<br />

L. monocy<strong>to</strong>genes has adapted its chitinases <strong>to</strong> exploit recognition <strong>of</strong> host substrates,<br />

potentially carbohydrate-linked molecules, <strong>to</strong> enhance bacterial survival<br />

within infected animals.<br />

Sub-lethal concentrations <strong>of</strong> common disinfectants do not influence<br />

survival and growth <strong>of</strong> Listeria monocy<strong>to</strong>genes in whole blood<br />

Holch, A., Gaedt Kastbjerg, V. and Gram, L.<br />

Technical University <strong>of</strong> Denmark<br />

Listeria monocy<strong>to</strong>genes is a serious food borne bacterial pathogen that can colonize<br />

food processing environment. Virulence gene expression is influenced by several environmental<br />

fac<strong>to</strong>rs e.g. temperature and oxygen-concentration. It has recently been<br />

shown that sub-lethal concentrations <strong>of</strong> disinfectants used in the food industry affect<br />

the expression <strong>of</strong> virulence genes in L. monocy<strong>to</strong>genes (Kastbjerg et al. 2009). The<br />

expression is either enhanced or reduced depending on the active compound in the<br />

disinfectant.The purpose <strong>of</strong> the study was <strong>to</strong> determine if these changes in virulence<br />

gene expression influence the virulence potential <strong>of</strong> L. monocy<strong>to</strong>genes in a more complex<br />

biological assay. We chose <strong>to</strong> study the ability <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> grow and<br />

survive in whole blood as several <strong>of</strong> the virulence genes are important for survival<br />

and growth in blood. Two strains <strong>of</strong> L. monocy<strong>to</strong>genes (EGD and a matern<strong>of</strong>etal strain)<br />

w<strong>here</strong> grown <strong>to</strong> exponential phase and exposed <strong>to</strong> two different disinfectants in a<br />

non-inhibi<strong>to</strong>ry and a sub-lethal concentration for one hour. The disinfectants contained<br />

either peroxide or quaternary ammonium compound (QAC) as their active<br />

compound. These cause enhanced or reduced virulence gene expression, respectively.<br />

The exposed bacteria were mixed with whole blood and the survival and growth was<br />

followed for 50 hours by plate counting. L. monocy<strong>to</strong>genes was able <strong>to</strong> grow in whole<br />

blood after exposure <strong>to</strong> the two different disinfectants in two different concentrations.<br />

However, pre-exposure <strong>to</strong> peroxide or QAC did not cause any difference in<br />

growth for both strains <strong>of</strong> L. monocy<strong>to</strong>genes as compared <strong>to</strong> exposure <strong>to</strong> water (control).<br />

In all experiments, the matern<strong>of</strong>etal strain grew slightly better than the EGD<br />

strain. Hence, the effect <strong>of</strong> sub-lethal concentrations <strong>of</strong> disinfectants on virulence<br />

gene expression did not affect the survival and growth in a complex biological model.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Internalin LRR domain variability in Listeria monocy<strong>to</strong>genes isolated<br />

from different hosts<br />

Zaytseva, E. A. 1,2, Ermolaeva, S. A. 2 and Somov, G. P. 1<br />

1. Research Institute for Epidemiology and Microbiology, RAMS Siberian Division (Vladivos<strong>to</strong>k)<br />

2. N.F. Gamaleya Research Institute for Epidemiology and Microbiology, RAMS (Moscow)<br />

Development <strong>of</strong> certain clinic manifestations <strong>of</strong> listeriosis has been recently suggested<br />

<strong>to</strong> depend on several fac<strong>to</strong>rs <strong>of</strong> L. monocy<strong>to</strong>genes pathogenicity. A number<br />

<strong>of</strong> surface and secreted proteins from internalin family, <strong>of</strong> which A and B internalin<br />

proteins are most thoroughly investigated, are involved in the process <strong>of</strong> Listeria<br />

invasion in<strong>to</strong> eukaryotic cells. One <strong>of</strong> the specific features <strong>of</strong> this protein family<br />

is presence <strong>of</strong> so called LLR domains involved in the direct interaction with<br />

eukaryotic recep<strong>to</strong>rs. The purpose <strong>of</strong> this research is <strong>to</strong> analyze distribution <strong>of</strong><br />

various internalin gene alleles among L. monocy<strong>to</strong>genes isolated from different<br />

sources. Eighty-six L. monocy<strong>to</strong>genes cultures were used belonging <strong>to</strong> serovariants<br />

1/2a, 1/2 b, 4b, isolated from various sources in the Far East and European part <strong>of</strong><br />

Russia in the period from 1952 <strong>to</strong> 2005. Primers, gene restricting fragment and<br />

coding LRR domain were selected with the help <strong>of</strong> Oligo38 application. Gene presence<br />

in L. monocy<strong>to</strong>genes was determined with the help <strong>of</strong> PCR. DNA fragment sequencing<br />

was performed in Genom Center (Moscow). Molecular genetics features<br />

<strong>of</strong> Listeria were analyzed in 4 culture groups depending on isolation source: 1)<br />

stillborn who died from listerial infection (n=21); 2) wild murine rodents (n=19); 3)<br />

marine hydrobionts (n=19); 4) food (control, n=27). Invasion fac<strong>to</strong>r genes - inlA,<br />

inlB, inlC and inlE - were found in all <strong>of</strong> L. monocy<strong>to</strong>genes isolates. Sequence <strong>of</strong><br />

LRR domains <strong>of</strong> genes inlA, inlB, inlC and inlE was determined for all the isolates.<br />

L. monocy<strong>to</strong>genes isolates demonstrated specific distribution <strong>of</strong> the examined gene<br />

alleles depending from isolation source. Among clinic isolates (group 1) low variability<br />

was recorded for inlA and inlC (p < 0.01). L. monocy<strong>to</strong>genes isolates obtained<br />

from wild rodent organs showed low variability with for inlB (p < 0.01). Specific<br />

alleles <strong>of</strong> inlA, inlC and inlE were found in Listeria cultures isolated from marine<br />

hydrobionts. Alleles <strong>of</strong> inlE were uniform in all the compared groups. In control<br />

group all gene alleles were distributed uniformly <strong>to</strong>o. The obtained results confirm<br />

internalin role in L. monocy<strong>to</strong>genes interaction with a certain host. A microbe<br />

with specific gene allele may be more virulent and invasive for a certain type <strong>of</strong><br />

host even at low infection doses.<br />

Reduced virulence <strong>of</strong> an adenylosuccinate lyase transposon mutant<br />

<strong>of</strong> a serotype 4b strain <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Faith, N. G. 1,2, Kim, J.-W. 3, Kathariou, S. 3, Sahaghian, R. 1 and Luchansky, J. B. 4<br />

1. School <strong>of</strong> Veterinary Medicine, Univ. Wisconsin-Madison Madison, WI<br />

2. Food Research Institute, Univ. Wisconsin-Madison Madison, WI<br />

3. Department <strong>of</strong> Food, Bioprocessing and Nutrition Sciences, North Carolina State Univ., Raleigh, NC<br />

4. Eastern Regional Research Labora<strong>to</strong>ry, USDA-ARS, Wyndmoor, PA<br />

A severe outbreak <strong>of</strong> listeriosis occurred in 1998–99 as a result <strong>of</strong> contamination <strong>of</strong><br />

hot dogs with serotype 4b Listeria monocy<strong>to</strong>genes. We compared several characteristics<br />

<strong>of</strong> strain H7550, implicated in the 1998–99 outbreak <strong>of</strong> listeriosis, and a plasmid-free<br />

derivative (H7550cds) <strong>of</strong> that strain lacking the cadmium resistance plasmid<br />

pLM80, with that <strong>of</strong> selected transposon mutant derivatives <strong>of</strong> those strains.<br />

We assessed virulence in a mouse model <strong>of</strong> intragastric (i.g.) inoculation <strong>of</strong> anesthetized<br />

A/J mice with approximately 10 6 CFU <strong>of</strong> the individual strains. A transposon<br />

mutant <strong>of</strong> strain H7550 that lacks adenylosuccinate lyase activity (strain<br />

J22F), and a non-hemolytic (LLO-) transposon mutant <strong>of</strong> strain H7550cds (J29H),<br />

were avirulent in our mouse model. We did not recover viable cells from the spleen,<br />

liver, blood gallbladder, or ceca <strong>of</strong> mice inoculated with either strain J22F or J29H,<br />

w<strong>here</strong>as the respective parent strains H7550 and its cadmium sensitive derivative<br />

H7550cds were equally virulent for mice. We observed no significant difference in<br />

the resistance <strong>of</strong> stationary phase cells <strong>of</strong> the plasmid harboring versus plasmidfree<br />

strains <strong>to</strong> synthetic gastric fluid at pH 4.5. All four strains formed bi<strong>of</strong>ilms on<br />

plastic surfaces within 24 hr in vitro. Strain J22F was better able <strong>to</strong> form bi<strong>of</strong>ilms in<br />

vitro than its parent strain H7550, w<strong>here</strong>as the LLO-negative mutant strain J29H<br />

was not significantly different from its parent strain H7550cds in bi<strong>of</strong>ilm formation.<br />

These results provide the first evidence that adenylosuccinate lyase activity is required<br />

for virulence <strong>of</strong> L. monocy<strong>to</strong>genes in mice. They also demonstrate that LLO<br />

is not required for bi<strong>of</strong>ilm formation in vitro.<br />

R E F E R E N C E<br />

B / P<br />

73<br />

R E F E R E N C E<br />

B / P<br />

74


R E F E R E N C E<br />

B / P<br />

75<br />

R E F E R E N C E<br />

B / P<br />

76<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Manifestations and outcome <strong>of</strong> listeriosis in adult patients<br />

Fernández Guerrero, M L. 1, Mancebo Plaza, B. 2, Torres, R. 2, Górgolas, M. 1 and Jusdado, J. J. 2<br />

1. Fundación Jiménez Díaz. Universidad Autónoma de Madrid, Spain<br />

2. Hospital Severo Ochoa, Madrid, Spain<br />

Infections caused by Listeria monocy<strong>to</strong>genes in adult patients are rarely reported<br />

despite the increasing prevalence <strong>of</strong> these infections in western countries. Hence,<br />

the manifestations, optimal treatment, prognosis and risk fac<strong>to</strong>rs for mortality<br />

are not entirely known. We <strong>here</strong>in report a review <strong>of</strong> the most relevant clinical aspects,<br />

treatment and outcome <strong>of</strong> a large series <strong>of</strong> patients with listeriosis studied in<br />

two hospitals in Madrid, over a period <strong>of</strong> 15 years. Patient with isolation <strong>of</strong> L. monocy<strong>to</strong>genes<br />

from blood, CSF or other sterile fluids were included in the analysis.<br />

Sixty-two cases were seen. The mean age was 58.5 years (from 20 <strong>to</strong> 85 years) without<br />

differences in distribution by genders. Seventy percent <strong>of</strong> cases had comorbidities:<br />

cirrhosis <strong>of</strong> the liver, hema<strong>to</strong>logic neoplasms, immunosupressive disorders<br />

and corticosteroid therapy were the most common. Primary bacteremia<br />

(48 %) and meningoencephalitis (37 %) were the most common presentations. Ten<br />

patients (16 %) presented with focal infections such as bacterial peri<strong>to</strong>nitis (4),<br />

endocarditis (3) and brain abscesses (3). Ampicillin alone or in combination with<br />

gentamicin and cotrimoxazol were the antimicrobial treatments most frequently<br />

used. We did not find differences in the mortality rate among patients treated with<br />

monotherapy in comparison with those that received combined therapy. Cotrimoxazol<br />

was successfully used in 10 out <strong>of</strong> 13 patients (77 %) treated. Twenty-three<br />

patients (37 %) died within 30-days <strong>of</strong> diagnosis. Mortality increased according <strong>to</strong><br />

the severity <strong>of</strong> the underlying disease and was significantly higher in immunocompromised<br />

patients (66 %) than in those with chronic debilitating conditions<br />

(20 %; p < 0.001). Human listeriosis is a severe disease producing high mortality<br />

particularly in immunocompromised patients with hema<strong>to</strong>logic neoplasms and<br />

corticosteroid therapy. Despite synergistic activity <strong>of</strong> combined ampicillin/gentamicin<br />

therapy, the survival benefits <strong>of</strong> this combination remain theoretical. Cotrimoxazol<br />

may be a useful alternative treatment for human listeriosis.<br />

Model for human Listeriosis: in vivo moni<strong>to</strong>ring <strong>of</strong> orally infected mice<br />

using bioluminescent Listeria monocy<strong>to</strong>genes<br />

Bergmann, S., Lengeling, A., Pasche, B. and Schughart, K.<br />

Helmholtz-Centre for Infectious Research, Germany<br />

The ubiqui<strong>to</strong>us Gram-positive bacterium Listeria monocy<strong>to</strong>genes is the causative<br />

agent <strong>of</strong> listeriosis. After ingestion <strong>of</strong> contaminated food, this pathogen is able <strong>to</strong><br />

cross the intestinal, blood-brain and placental barrier and leads <strong>to</strong> gastroenteritis,<br />

meningitis and matern<strong>of</strong>etal infections which may result in abortion and spontaneous<br />

stillbirth. We have generated a bioluminescent Listeria monocy<strong>to</strong>genes strain<br />

which carries two amino acid substitutions in the bacterial invasion protein internalin<br />

A (InlA S192NY369S) and enables the bacterium <strong>to</strong> bind <strong>to</strong> the murine Ecadherin<br />

recep<strong>to</strong>r with increased affinity as compared <strong>to</strong> wildtype Listeria. The<br />

genetic modification allows the bacteria <strong>to</strong> invade intestinal epithelial cells and<br />

<strong>to</strong> cross the murine intestinal barrier with high efficiency. The new bioluminescent<br />

Listeria strain was used <strong>to</strong> moni<strong>to</strong>r bacterial dissemination in orally infected mice.<br />

For visualization <strong>of</strong> bioluminescent bacteria during infection we have used the<br />

Xenogen IVIS system, a highly sensitive, low light system optimized for in vivo imaging.<br />

Here, we present first results <strong>of</strong> different mouse inbred strains (BALB/cByJ,<br />

C57BL/6J, CD1, 129P, C3HeB/FeJ) that behave different in the intensity <strong>of</strong> infection<br />

and show a time variation regarding bacterial spreading, the crisis <strong>of</strong> infection<br />

and bacterial clearance. We found CD1 and C3HeB/FeJ mice <strong>to</strong> be resistant <strong>to</strong><br />

orally transmitted listeriosis, they showed reduced bacterial dissemination in the<br />

early phase <strong>of</strong> infection (during first 72 hrs), and a fast clearance <strong>of</strong> the pathogen<br />

from day 5 p.i. on. In contrast BALB/cByJ and C57BL/6J mice we found <strong>to</strong> be more<br />

susceptible after oral infection. This is reflected by slower bacterial clearance and<br />

a reduced survival.<br />

We further give an outlook how we will use this new approach <strong>to</strong> obtain a detailed<br />

insight in the process <strong>of</strong> crossing the fe<strong>to</strong>-placental barrier in pregnant mice as<br />

well as the blood-brain barrier after oral infection with our modified Listeria<br />

monocy<strong>to</strong>genes strain.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Galleria mellonella as model system <strong>to</strong> study Listeria species-specific<br />

and Listeria monocy<strong>to</strong>genes serotype-specific pathogenesis<br />

Mraheil, M. A., Krishnendu, M., Hain, T. and Chakraborty, T.<br />

Germany<br />

Essential aspects <strong>of</strong> the innate immune response <strong>to</strong> microbial infection are conserved<br />

between insects and mammals. This has generated interest in using insects<br />

as model organisms <strong>to</strong> study host-microbe interactions. We used the greater wax<br />

moth Galleria mellonella, which can be reared at 37 °C, as a model host for examining<br />

virulence potential <strong>of</strong> Listeria spp. Here we report that Galleria is an excellent<br />

surrogate model <strong>of</strong> listerial septic infection, capable <strong>of</strong> clearly distinguishing<br />

between pathogenic and non-pathogenic Listeria and even between virulent and<br />

attenuated Listeria monocy<strong>to</strong>genes strains and seotypes. Virulence required listerial<br />

genes hither<strong>to</strong> implicated in the mouse infection model, and was linked <strong>to</strong><br />

strong antimicrobial activities in both hemolymph and hemocytes <strong>of</strong> infected larvae.<br />

We conclude that severity <strong>of</strong> septic infection <strong>of</strong> L. monocy<strong>to</strong>genes is primarily<br />

modulated by innate immune responses and suggest the use <strong>of</strong> Galleria as a relatively<br />

simple, non-mammalian model system that can be used <strong>to</strong> assess the virulence<br />

<strong>of</strong> strains <strong>of</strong> Listeria spp. isolated from a wide variety <strong>of</strong> settings from both<br />

the clinic and the environment.<br />

Listeria monocy<strong>to</strong>genes ActA is a key player<br />

in evading au<strong>to</strong>phagic recognition<br />

Pillich, H., Loose, M., Hain, T. and Chakraborty, T.<br />

Germany<br />

Au<strong>to</strong>phagy is a pivotal bulk degradation system that eliminates undesirable molecules,<br />

damaged organelles, and misfolded protein aggregates in response <strong>to</strong> diverse<br />

stimuli, including infection. Au<strong>to</strong>phagy acts <strong>to</strong> limit intracellular microbial<br />

growth but intracellular pathogens have evolved strategies <strong>to</strong> subvert host au<strong>to</strong>phagic<br />

responses for their survival. We found that Listeria monocy<strong>to</strong>genes ActA,<br />

a surface protein required for actin polymerization and actin-based bacterial<br />

motility, plays a pivotal role in evading au<strong>to</strong>phagy, but in a manner independent <strong>of</strong><br />

bacterial motility. We show that L. monocy<strong>to</strong>genes exploits the biomimetic property<br />

<strong>of</strong> ActA <strong>to</strong> camouflage itself with host proteins comprised <strong>of</strong> Ena/VASP and<br />

the Arp2/3 complex, t<strong>here</strong>by escaping recognition by au<strong>to</strong>phagy.<br />

R E F E R E N C E<br />

B / P<br />

77<br />

R E F E R E N C E<br />

B / P<br />

78


R E F E R E N C E<br />

B / P<br />

79<br />

R E F E R E N C E<br />

B / P<br />

80<br />

AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

Non-haemolytic and hypovirulent Listeria monocy<strong>to</strong>genes became<br />

haemolytic and virulent after passage through mice<br />

Secic, I., Lindbäck, T. and Rørvik, L. M.<br />

Norwegian School <strong>of</strong> Veterinary Science<br />

Two different isolates (from a fish processing plant and a broiler abat<strong>to</strong>ir) showing<br />

typical appearance for L. monocy<strong>to</strong>genes on OCLA and ALOA, did not express<br />

haemolysis on blood agar with blood from different sources (sheep, horse, bovine,<br />

human) and appeared as L. innocua on RLM. However, biochemical reactions and<br />

16S RNA analysis identified them as L. monocy<strong>to</strong>genes. High levels (10 9) <strong>of</strong> each<br />

isolate were injected i.p. in<strong>to</strong> five mice. Three <strong>of</strong> the mice died after 1-3 days for<br />

both strains. Cultivation from livers and spleens revealed a mixture <strong>of</strong> haemolytic<br />

and non-haemolytic colonies. Non-haemolytic isolates showed typical appearance<br />

for L. monocy<strong>to</strong>genes on OCLA and ALOA, but were identified as L. innocua on<br />

RLM. Haemolytic and non-haemolytic isolates from each mouse, as well as the respective<br />

mother strains were identical pulsotypes. In contrast <strong>to</strong> passage through<br />

mice, fifty passages <strong>of</strong> the mother strains on RLM did not change the phenotype,<br />

and screening <strong>of</strong> 10 7 colonies did not reveal any haemolytic subpopulation. HT-<br />

29 cell assay showed that while the non-haemolytic isolates were all hypovirulent,<br />

their haemolytic twins were virulent. Part <strong>of</strong> the LIPI 1 virulence island <strong>of</strong> nonhaemolytic<br />

and haemolytic isolates was sequenced. The non-haemolytic isolate<br />

had a seven base insertion in the prfA gene which introduced a s<strong>to</strong>p codon, resulting<br />

in a truncated PrfA protein. The seven base pair insertion was deleted during<br />

infection in mice, resulting in a normal PrfA protein in the haemolytic version.<br />

This study shows that during passage through mice hypovirulent L. monocy<strong>to</strong>genes<br />

can change <strong>to</strong> a virulent phenotype. It is not known if this phenomenon may occur<br />

in humans. These strains would not have been recognized as L. monocy<strong>to</strong>genes unless<br />

OCLA or ALOA had been used as isolation agar.<br />

Mutants <strong>of</strong> Listeria monocy<strong>to</strong>genes (Lm) resistant <strong>to</strong> the polycationic<br />

peptide protamine appear <strong>to</strong> be attenuated for virulence<br />

Schlech, W.<br />

Dalhousie University, Canada<br />

Previously, we reported the isolation <strong>of</strong> Lm mutants resistant <strong>to</strong> protamine, a relatively<br />

inexpensive polycationic peptide derived from salmon and herring milt.<br />

Here, we report the final characterization <strong>of</strong> two protamine-resistant mutants, emphasizing<br />

their virulence defects, which have led us <strong>to</strong> believe that protamine-resistant<br />

mutants are attenuated. We also report that p60 mutants, isolated independently<br />

from the protamine-resistance phenotype and previously reported <strong>to</strong> show<br />

virulence defects, are also resistant <strong>to</strong> protamine. Protamine-resistant (PtmR) mutants<br />

from the Canadian Maritimes outbreak strain 15U and from the Massachusetts<br />

outbreak strain Scott A, were picked as isolated colonies from TSA plates containing<br />

1 mg/ml <strong>of</strong> protamine. Electron microscopy and SDS-PAGE were<br />

extensively applied <strong>to</strong> the initial characterization <strong>of</strong> the mutants. Swiss-Webster<br />

mice were infected by direct gastric inoculation with different Lm doses and spread<br />

<strong>to</strong> the spleen and liver moni<strong>to</strong>red at 3 days after inoculation. Survival <strong>of</strong> Lm in diffusion<br />

chambers surgically implanted in the peri<strong>to</strong>neal cavity <strong>of</strong> rats for 3 days, was<br />

also evaluated. By electron microscopy and SDS-PAGE, the PtmR mutants displayed<br />

significant differences in their surface proteins and structure. However,<br />

both showed a reduced ability <strong>to</strong> infect mice (particularly the liver) and a reduced<br />

survival in the intraperi<strong>to</strong>neal diffusion chambers. p60 mutants, which are unable<br />

<strong>to</strong> efficiently infect cells in culture, turned out <strong>to</strong> also be PtmR. It seems that PtmR<br />

mutants, in spite <strong>of</strong> their different origin and surface characteristics, are attenuated<br />

for virulence. This possibly represents an example <strong>of</strong> an inverse fitness-virulence<br />

relationship that could be exploited <strong>to</strong> select against highly virulent Lm strains in<br />

food products treated with inhibi<strong>to</strong>ry concentrations <strong>of</strong> polycationic peptides.


AREA // B // Listeria monocy<strong>to</strong>genes as a human and animal pathogen<br />

POSTER PRESENTATIONS // P / 47–81, 182<br />

The role <strong>of</strong> plasmacy<strong>to</strong>id dendritic cells in the course <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

infection<br />

Solodova, E., Lienenklaus, S., Jablonska, J. and Weiss, S.<br />

Labora<strong>to</strong>ry <strong>of</strong> Molecular Immunology, Helmholtz Centre for Infection Research, Braunschweig, Germany<br />

Type I interferons (IFNs) play a key role in linking the innate and adaptive arms <strong>of</strong><br />

the immune system. T<strong>here</strong> are more that 13 IFN-α and a single IFN-β, all using a<br />

common recep<strong>to</strong>r – IFNAR, which is expressed on wide variety <strong>of</strong> cell types. Production<br />

<strong>of</strong> type I IFNs in response <strong>to</strong> infection with viruses is essential for clearance<br />

<strong>of</strong> the pathogen from the host. But the impact <strong>of</strong> these cy<strong>to</strong>kines during bacterial infection<br />

is less defined. Plasmacy<strong>to</strong>id dendritic cells (pDCs) are known <strong>to</strong> be major<br />

natural type I IFN producing cells during viral infections, but not much is known<br />

about their impact in the course <strong>of</strong> bacterial infections. Listeria monocy<strong>to</strong>genes is<br />

one <strong>of</strong> the bacteria known <strong>to</strong> induce type I IFN synthesis. In contrast <strong>to</strong> viral infections<br />

these cy<strong>to</strong>kines were shown <strong>to</strong> have detrimental effects for the host during<br />

infection with this bacterium. In our study we used tissue specific conditional reporter<br />

and knock-out mice showing that LysMcre-expressing cells but not pDCs<br />

are responsible for type I IFN production during Listeria monocy<strong>to</strong>genes infection.<br />

We have also shown that depletion <strong>of</strong> pDCs revealed a phenotype showing the ability<br />

<strong>of</strong> pDCs <strong>to</strong> protect mice infected with Listeria monocy<strong>to</strong>genes. The aim <strong>of</strong> my<br />

present work is <strong>to</strong> investigate this phenomenon und <strong>to</strong> understand how pDCs are<br />

able <strong>to</strong> provide protection <strong>to</strong> mice during Listeria monocy<strong>to</strong>genes infection.<br />

Oxygen restriction increases the infection potential<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes – a transcriptional analysis<br />

Andersen, J. B., Bergstrøm, A., Knudsen, G., Bak Christensen, B., Ebersbach, T.,<br />

Boye, M. and Rask Licht, T.<br />

Technical University <strong>of</strong> Denmark, National Food Institute, DTU, Division <strong>of</strong> Microbiology and Risk<br />

Assessment, Denmark<br />

Listeria monocy<strong>to</strong>genes has been implicated in several food borne outbreaks as<br />

well as sporadic cases <strong>of</strong> disease during the last two decades. Increased understanding<br />

<strong>of</strong> the biology <strong>of</strong> this organism is important in the prevention <strong>of</strong> food<br />

borne listeriosis. This is highly relevant for safety assessment <strong>of</strong> this organism in<br />

food. We have previously shown (Andersen et al., BMC Microbiology; 2007, 7:55)<br />

that the environmental conditions <strong>to</strong> which L. monocy<strong>to</strong>genes is exposed prior <strong>to</strong><br />

ingestion are decisive for its in vivo infective potential in the gastrointestinal tract<br />

after passage <strong>of</strong> the gastric barrier. Infection <strong>of</strong> Caco-2 cells revealed that Listeria<br />

cultivated under oxygen-restricted conditions were approximately 100 fold more<br />

invasive than similar cultures grown without oxygen restriction. This means that<br />

not only the number <strong>of</strong> Listeria present in a given food item, but that also the physiological<br />

condition <strong>of</strong> these bacteria is important for food safety. The in vitro and in<br />

vivo data suggest that an oxygen-restricted L. monocy<strong>to</strong>genes cell represents a significantly<br />

higher risk than a cell grown without oxygen restriction. In order <strong>to</strong><br />

identify transcriptional differences contributing <strong>to</strong> different invasiveness, microarray<br />

gene chip technology was applied <strong>to</strong> cDNA created from RNA isolated<br />

from oxygen restricted and non-restricted cultures. The analysis confirmed several<br />

relevant genes <strong>to</strong> be differentially transcribed in the two environmental conditions<br />

e.g. genes related <strong>to</strong> virulence potential <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

R E F E R E N C E<br />

B / P<br />

81<br />

R E F E R E N C E<br />

B / P<br />

182


R E F E R E N C E<br />

C / P<br />

82<br />

R E F E R E N C E<br />

C / P<br />

83<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Semi-au<strong>to</strong>mated repetitive sequenced-based PCR compared <strong>to</strong> pulsed<br />

field gel electrophoresis for Listeria monocy<strong>to</strong>genes sub-typing<br />

Roussel, S., Félix, B., Vignaud, M.-L., Tam Dao, T., Marault, M. and Brisabois, A.<br />

AFSSA, France<br />

Listeriosis is a severe infection which mainly impacts pregnant women, neonates<br />

and immuno-compromised adults. The commercially available, semi-au<strong>to</strong>mated<br />

rep-PCR assay system, DiversiLab, has been successfully used for subtyping several<br />

species <strong>of</strong> bacteria. We compared <strong>here</strong> the DiversiLab System with macrorestriction<br />

analysis by pulsed field gel electrophoresis (PFGE) which is currently the gold<br />

standard for molecular subtyping <strong>of</strong> Listeria monocy<strong>to</strong>genes. We used a panel <strong>of</strong><br />

116 human and food L. monocy<strong>to</strong>genes strains for the comparative evaluation.<br />

Among these strains, t<strong>here</strong> were 4 pairs <strong>of</strong> duplicates; 13 strains were epidemiologically<br />

related and the remaining food isolates were epidemiologically unrelated.<br />

The strains <strong>of</strong> different serotypes revealed distinct ApaI-PFGE types, Rep-PCR<br />

types (RT) and AscI- PFGE types except for one RT and one AscI- PFGE type. The<br />

four doublets displayed a unique RT, AscI and ApaI PFGE pattern showing the<br />

good reproducibility <strong>of</strong> the three methods. The epidemiologically related strains<br />

were clustered in the same RT and PFGE types. The Simpson’s index <strong>of</strong> diversity<br />

was 0.954; 0.988; 0.994; 0.998 for DiversiLab, AscI-PFGE, ApaI-PFGE and<br />

AscI/ApaI-PFGE respectively. Thus, PFGE was more discriminating than DiversiLab.<br />

However, for 1/2a serotype strains, six AscI-PFGE and three ApaI-PFGE<br />

types were divided in<strong>to</strong> different RT. DiversiLab allowed a good discrimination<br />

between serotype 1/2a strains. This investigation also demonstrated the ability <strong>of</strong><br />

the DiversiLab <strong>to</strong> differentiate serotypes 4b and 1/2b strains. DiversiLab is less<br />

labour-intensive than PFGE and provides results in less than 24 hours compared<br />

with 30 hours <strong>to</strong> 3 days for PFGE from the time a pure culture <strong>of</strong> the bacteria is obtained.<br />

Based on these results, DiversiLab may be useful for tracking the source <strong>of</strong><br />

contamination in food processing facilities and their environments.<br />

Listeriosis: a frequent cause <strong>of</strong> fatal encephalitis in France<br />

with high case fatality<br />

Mailles, A. 1, Vaillant, V., Lecuit, M. 2 and Stahl, J.-P. 3<br />

1. Institut de Veille Sanitaire, France<br />

2. Institut Pasteur, France<br />

3. University Hospital <strong>of</strong> Grenoble, France<br />

In 2007, we carried out a national prospective study in 106 hospital units <strong>to</strong> assess<br />

the aetiology, clinical patterns and outcome <strong>of</strong> infectious encephalitis in<br />

France. Among 253 case-patients, encephalitis due <strong>to</strong> L. monocy<strong>to</strong>genes was identified<br />

in 13 people. A case <strong>of</strong> encephalitis was a patient aged at least 28 days, hospitalised<br />

in France with an acute onset <strong>of</strong> illness and at least one abnormality <strong>of</strong><br />

the CSF, and fever and decreased consciousness or seizures or altered mental status<br />

or focal neurological signs. All case-patients had a complete biological investigation<br />

<strong>to</strong> assess the etiologic diagnosis. Clinical data were collected using standardised<br />

questionnaires. Listeriosis case-patients were compared <strong>to</strong> other<br />

case-patients enrolled in the study. Thirteen case-patients were identified with<br />

encephalitis due <strong>to</strong> L. monocy<strong>to</strong>genes among 253 case-patients (5 %) and 131 casepatients<br />

with an etiological diagnosis (10 %). Listeriosis was the fourth most frequent<br />

cause <strong>of</strong> encephalitis identified after HSV (n=55), VZV (n=20) and tuberculosis<br />

(n = 20). Listeria was isolated in CSF from XX patients, from blood only from<br />

1 case. XX cases had a positive PCR on CSF. The last patient was diagnosed on a<br />

combination <strong>of</strong> clinical and epidemiological criteria. Eleven listeriosis case-patients<br />

were confirmed cases, 1 was a probable case and 1 was a possible case. Their<br />

mean age was 71 years and 9 were men. Listeriosis case-patients were more likely<br />

<strong>to</strong> have a current treated cancer (23 % vs 5 %, p=0.005) and <strong>to</strong> present with cranial<br />

nerves impairments (38 % vs 17 %, p=0.02). Six <strong>of</strong> 13 (46 %) listeriosis casepatients<br />

had a fatal outcome vs 20/216 (8.5 %) <strong>of</strong> other patients (p=0.007). These<br />

results showed that L. monocy<strong>to</strong>genes is a frequent cause <strong>of</strong> encephalitis in France<br />

and confirmed its severity. Listeriosis should be considered early in the course <strong>of</strong><br />

encephalitis, as specific and efficient treatment can be proposed.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Listeria monocy<strong>to</strong>genes: identification and subtyping<br />

Favretti, M.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale delle Venezie, Italy<br />

Listeria monocy<strong>to</strong>genes is a common environmental organism and an important<br />

food-borne pathogen <strong>to</strong>o. Pregnant women, neonates, elderly and immunocompromised<br />

patients are at greatest risk <strong>of</strong> acquiring listeriosis. Since the recognition<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes as a food-borne pathogen t<strong>here</strong> have been rapid advances<br />

in the development <strong>of</strong> suitable methods for isolation and identification.<br />

Although all Listeria monocy<strong>to</strong>genes strains are considered pathogen, only some<br />

serotypes are predominant and frequently involved in outbreaks. So it is very important<br />

<strong>to</strong> obtain viable isolates for typing, elucidating outbreaks and tracing the<br />

source <strong>of</strong> infection. The differentiation <strong>of</strong> Listeria monocy<strong>to</strong>genes isolates with<br />

phenotypic or molecular subtyping methods has provided useful information related<br />

<strong>to</strong> the phylogeny, epidemiology and ecology <strong>of</strong> the organism and also has permitted<br />

the correlation <strong>of</strong> isolates from food and from human case in order <strong>to</strong> underline<br />

any relationship between them. Although t<strong>here</strong> are several techniques for<br />

identification and typing available at the moment, it is not possible <strong>to</strong> identify the<br />

“gold standard” method yet. T<strong>here</strong>fore, using in parallel different methods ensures<br />

a correct and complete identification <strong>of</strong> strains. We present an overview <strong>of</strong><br />

the techniques used <strong>to</strong> isolate, <strong>to</strong> identify and <strong>to</strong> subtype Listeria monocy<strong>to</strong>genes;<br />

the various subtyping systems provide different degrees <strong>of</strong> discrimination among<br />

isolates. Subtyping methods for Listeria monocy<strong>to</strong>genes include phenotypic (e.g.<br />

serotyping and phage-typing) and different DNA–based subtyping methods (e.g.<br />

multilocus enzyme electrophoresis, ribotyping, pulsed-field gel electrophoresis<br />

(PFGE), polymerase chain reaction (PCR)).<br />

Virulotyping <strong>of</strong> Listeria monocy<strong>to</strong>genes by high resolution melt analysis<br />

Amar, C. F. 1, Tamburro, M. 2, Dear, P. 3 and Grant, K. 1<br />

1. Health Protection Agency, Centre for Infections, UK<br />

2. University <strong>of</strong> Molise, Campobasso, Italy<br />

3. Labora<strong>to</strong>ry <strong>of</strong> Molecular Biology, Cambridge, UK<br />

L. monocy<strong>to</strong>genes causes a severe foodborne infection in vulnerable people. Isolates<br />

are known <strong>to</strong> vary in their ability <strong>to</strong> cause infection but current typing methods<br />

do not provide information on strain pathogenicity. Six key L. monocy<strong>to</strong>genes<br />

virulence genes, clustered <strong>to</strong>gether on pathogenicity island, LIPI-1 (9Kb), are<br />

known <strong>to</strong> be essential for intracellular survival, whilst two other virulence genes,<br />

internalin A and B, are responsible for host cell invasion. Many food isolates produce<br />

truncated internalin proteins due <strong>to</strong> point mutations and thus have reduced<br />

pathogenicity. Detection <strong>of</strong> these mutations and those in other virulence genes<br />

are likely <strong>to</strong> provide valuable information on strain pathogenicity. High Resolution<br />

Melt (HRM) analysis is a powerful technique capable <strong>of</strong> identifying sequence variations<br />

in PCR amplicons by accurately determining their melting temperature<br />

(Tm). This technique was used <strong>to</strong> generate virulence fingerprints, using 81 markers<br />

spanning LIP1 and Internalin A and B genes. Strains <strong>of</strong> L. monocy<strong>to</strong>genes from<br />

human cases, food and the environment were compared with results generated<br />

using L. monocy<strong>to</strong>genes EGDe. Considerable variations in Tm were found across<br />

markers for LIP1 and internalin A and B genes which were confirmed by sequencing.<br />

Eight markers had Tms specific <strong>to</strong> genetic lineage I and 11 <strong>to</strong> genetic lineage II.<br />

Fifteen markers had conserved sequences across both lineages with those spanning<br />

the hlyA gene being the most conserved. In the housekeeping gene, prs,<br />

marker variations were only detected in serotype 4b strains. Tm variations were<br />

detected in markers for actA, intA and intB genes in which point mutations are<br />

associated with the expression <strong>of</strong> truncated proteins and a concomitant reduction<br />

in virulence. HRM is an accurate and rapid technique for detecting sequence diversity<br />

within bacterial genes and not only presents a novel method for strain characterisation<br />

but also <strong>of</strong>fers a unique potential <strong>to</strong> inform on strain pathogenicity.<br />

R E F E R E N C E<br />

C / P<br />

84<br />

R E F E R E N C E<br />

C / P<br />

85


R E F E R E N C E<br />

C / P<br />

86<br />

R E F E R E N C E<br />

C / P<br />

87<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Risk fac<strong>to</strong>rs for nonperinatal listeriosis mortality<br />

in Los Angeles County, California, 1992–2004<br />

Guevara, R. E.*, Mascola, L. and Sorvillo, F.<br />

County <strong>of</strong> Los Angeles Department <strong>of</strong> Public Health, USA<br />

Listeriosis is a relatively rare foodborne disease but has significant public health<br />

implications as it has a case-fatality rate <strong>of</strong> 20 % and it causes 28 % <strong>of</strong> all foodborne<br />

disease related deaths. Yet, data on the risk fac<strong>to</strong>rs for listeriosis mortality<br />

have been limited. A 13-year retrospective cohort study was performed <strong>to</strong> describe<br />

nonperinatal listeriosis mortality in Los Angeles County, California, during<br />

1992–2004. A nonperinatal listeriosis case was defined as a non-pregnant person<br />

>42 days old with a positive culture <strong>of</strong> Listeria monocy<strong>to</strong>genes and residence in<br />

Los Angeles County. Causal modeling based on prior knowledge and bivariable<br />

analyses was performed <strong>to</strong> determine a comprehensive unconditional multivariable<br />

logistic regression model with 29 main effect variables. With 281 nonperinatal<br />

listeriosis cases, multivariable analysis found non-hema<strong>to</strong>logical malignancy<br />

(odds ratio: 5.92, 95 % confidence interval: 1.85-18.9), alcoholism (4.63, 1.36-15.8),<br />

age ≥ 70 years (3.44, 1.50-7.87), steroid medication (3.34, 1.38-8.08), and kidney<br />

disease (2.94, 1.18-7.31) <strong>to</strong> be statistically significant risk fac<strong>to</strong>rs for mortality.<br />

Other listeriosis mortality risk fac<strong>to</strong>rs with adjusted odds ratios >1.5 included<br />

blood transfusion, asthma, black race-ethnicity, Asian race-ethnicity, antibiotics,<br />

hypertension, chemotherapy, and Hispanic race-ethnicity. Cases admitted with<br />

sepsis only had the highest mortality (23.7 %), while cases with meningitis only<br />

had the lowest mortality (3.13%). The findings <strong>of</strong> this study should highlight the<br />

importance <strong>of</strong> certain underlying risk fac<strong>to</strong>rs and their association with nonperinatal<br />

listeriosis-related deaths. (This study was published in Clinical Infectious<br />

Diseases 2009; 48:1507-15 ©2009 by the Infectious Diseases Society <strong>of</strong> America.<br />

http://www.journals.uchicago.edu/loi/cid).<br />

* Participation Supported by IUFoST.<br />

Molecular typing <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated from ovine sausage<br />

Nives, M. R. 1, Mele, P. 1, Parisi, A. 2, La<strong>to</strong>rre, L. 2, Virgilio, S. 1 and Tola, S. 1<br />

1. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale <strong>of</strong> Sardinia, Italy<br />

2. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale <strong>of</strong> Apulia and Basilicata Italy<br />

Sardinia, an island located in the middle <strong>of</strong> Mediterranean sea, has about four million<br />

<strong>of</strong> milking sarda sheep. In recent years in addition <strong>to</strong> milk and cheese production<br />

other products from sheep meat have been commercialized (ham and<br />

sausage). In this survey, between 2007 – 2008, five sampling were performed inside<br />

a fac<strong>to</strong>ry located in the northern side <strong>of</strong> Sardinia. We moni<strong>to</strong>red the different<br />

stages <strong>of</strong> sausage production for detecting the presence <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in food matrices, the equipment and the environment. Isolation and identification<br />

<strong>of</strong> L. monocy<strong>to</strong>genes were carried out according <strong>to</strong> ISO 11290-1/2:1996/98<br />

amendment 1-2004. The isolates were characterized using serotyping and three<br />

different molecular methods: PFGE, MLST, VNTR according <strong>to</strong> previous published<br />

pro<strong>to</strong>cols. Moreover five virulence genes (actA, hlyA, plcB, prfA, iap) were detected<br />

by PCR. L. monocy<strong>to</strong>genes was isolated in all the phases <strong>of</strong> the sausage production<br />

(raw minced meat, meat mixture after salt and spice addition, sausage after desiccation<br />

time, sausage during different seasoning periods) and from the floor drain.<br />

On the whole 21 out <strong>of</strong> 126 samples resulted contaminated. The L. monocy<strong>to</strong>genes<br />

count resulted in all the samples higher than 100 UFC/g (Regulation EC n°<br />

2073/05). All isolates resulted PCR-positive for the investigated virulence genes.<br />

PFGE, MLST and VNTR identified two different molecular pr<strong>of</strong>iles. Five isolates,<br />

marked as type A and belonging <strong>to</strong> the serotype 1/2a, were isolated just from one<br />

batch samples w<strong>here</strong>as the type B (n=16), belonging <strong>to</strong> the serotype 1/2b, was isolated<br />

from three different food batches and from the floor drain. Our results<br />

demonstrate an high occurrence <strong>of</strong> L. monocy<strong>to</strong>genes in the studied facility. So the<br />

sheep sausage production must be object <strong>of</strong> a careful and constant vigilance activity<br />

and the isolates should be characterized using molecular methods for timely<br />

moni<strong>to</strong>ring the contaminations and <strong>to</strong> play out the due prevention actions.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

A novel phage-PCR assay for the rapid detection <strong>of</strong> viable Listeria monocy<strong>to</strong>genes<br />

in food within 24 h<br />

Elemam, M. M.<br />

School <strong>of</strong> Biosciences, University <strong>of</strong> Nottingham, UK<br />

Listeria monocy<strong>to</strong>genes is a pathogen <strong>of</strong> public health significance causing occasional<br />

outbreaks and sporadic cases <strong>of</strong> foodborne illness. This microorganism is <strong>of</strong><br />

particular concern in the food industry due <strong>to</strong> its widespread distribution in nature,<br />

its ability <strong>to</strong> survive in a wide range <strong>of</strong> environmental conditions, and its ability<br />

<strong>to</strong> grow at refrigeration temperatures. The standard microbiological method<br />

for L. monocy<strong>to</strong>genes detection in foods (ISO 11290-1/ A1 - 2004 and NHS-2009) is<br />

costly and time consuming; require at least 6 days obtaining a results. The development<br />

<strong>of</strong> a rapid and reliable test capable <strong>of</strong> detection very low numbers <strong>of</strong> the organism<br />

in ready <strong>to</strong> eat products is <strong>of</strong> significant importance <strong>to</strong> the food industry.<br />

The Phage Amplification assay has been developed as a rapid method for the detection<br />

<strong>of</strong> L. monocy<strong>to</strong>genes. We have been using the broad host range phage A511<br />

<strong>to</strong> develop such a method for detection <strong>of</strong> Listeria monocy<strong>to</strong>genes in both cheese<br />

and s<strong>to</strong>mached food samples. Successful development <strong>of</strong> the assay requires a virucide<br />

<strong>to</strong> achieve differential inactivation <strong>of</strong> the phage without affecting the viability<br />

<strong>of</strong> the target cell <strong>to</strong> be detected. Several different chemicals were evaluated as potential<br />

virucides, but tea extracts were found <strong>to</strong> be the most effective virucidal<br />

agent. The efficacy <strong>of</strong> the new assay was tested using cheese samples and was <strong>to</strong><br />

shown <strong>to</strong> be able <strong>to</strong> detect low numbers <strong>of</strong> cells, in only 24 h in compared <strong>to</strong> the 6<br />

days <strong>to</strong> conventional culture methods. Because <strong>of</strong> the broad host range <strong>of</strong> the<br />

phage the identity <strong>of</strong> the cell detected must be confirmed using PCR amplification<br />

<strong>of</strong> signature sequences from the phage plaque and several different PCR<br />

strategies have been evaluated. The combined phage-PCR method would allow<br />

rapid screening for food products prior <strong>to</strong> release from the fac<strong>to</strong>ry.<br />

Perinatal listeriosis in Los Angeles County, California, 1992 – 2004<br />

Guevara, R. E.* and Mascola, L.<br />

County <strong>of</strong> Los Angeles Department <strong>of</strong> Public Health, USA<br />

In Los Angeles County (LAC), California, physicians and labora<strong>to</strong>ries are required<br />

<strong>to</strong> report all listeriosis <strong>to</strong> the LAC Department <strong>of</strong> Public Health. From this passive<br />

surveillance, a population-based epidemiologic study was performed <strong>to</strong> further describe<br />

perinatal listeriosis. Perinatal listeriosis was defined when Listeria monocy<strong>to</strong>genes<br />

was identified from a normally sterile site between a mother-fetus or<br />

mother-infant (infant age < 43 days old) pair. During 1992 – 2004, <strong>of</strong> 413 listeriosis<br />

cases reported, 122 (30 %) were perinatal listeriosis cases. Perinatal listeriosis had<br />

a higher average annual incidence rate (5.9 cases/100,000 live births/year) and<br />

mortality (32.2 %) than nonperinatal listeriosis (2.5 cases/million people/year, and<br />

18.6 %, respectively). Disease onset during pregnancy defined intrapartum cases<br />

(n=93, 76 %), during 0-6 days after birth defined early onset cases (n=21, 17 %), and<br />

during 7-42 days after birth (n=8, 7 %) defined late onset cases. For intrapartum,<br />

early-onset, and late-onset cases, fetal/infant mortality was 36.6 % (n=34), 14.3%<br />

(n=3), and 0%, respectively. No maternal deaths occurred. Median gestational age at<br />

admission <strong>of</strong> intrapartum cases was 28 weeks (range 10-40 weeks) with maternal<br />

symp<strong>to</strong>ms commonly including fever (84 %), abdominal pain (35 %), chills (29 %),<br />

back pain (24 %), and headache (20 %). Intrapartum listeriosis cases had maternal<br />

infections <strong>of</strong> both sepsis and meningitis (n=2, 2.2 %) and sepsis only (n=72, 77.4 %),<br />

but mothers with symp<strong>to</strong>ms in 19 (20.4 %) intrapartum cases were culture-negative<br />

or not cultured. Among intrapartum cases with hospital admission for listeriosis<br />

before 38 weeks gestation, 33 (35 %) had fetuses continue in<strong>to</strong> pregnancy for<br />

more than 7 days (6 fetal demises/stillbirths with median <strong>of</strong> 13.5 days after onset, 4<br />

sick newborns with median 30.5 days after onset, 17 healthy newborns with median<br />

25 days after onset, and 6 unknown birth outcomes). Although annual incidence<br />

<strong>of</strong> perinatal listeriosis has been decreasing, opportunities such as culturing<br />

symp<strong>to</strong>matic mothers for L. monocy<strong>to</strong>genes exist <strong>to</strong> further prevent perinatal listeriosis<br />

morbidity and mortality.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

C / P<br />

88<br />

R E F E R E N C E<br />

C / P<br />

89


R E F E R E N C E<br />

C / P<br />

90<br />

R E F E R E N C E<br />

C / P<br />

91<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Antimicrobial resistance <strong>of</strong> Listeria monocy<strong>to</strong>genes human strains<br />

isolated since 1926 in France<br />

Morvan, C., Moubareck, A., Leclercq, M., Herve-Bazin, S., Bremont, M., Lecuit, P.<br />

and Le Monnier Courvalin, A.<br />

Institut Pasteur, France<br />

Amoxicillin plus gentamicin is the treatment <strong>of</strong> choice for listeriosis, a foodborne<br />

infection caused by Listeria monocy<strong>to</strong>genes (Lm). Lm is considered universally<br />

susceptible <strong>to</strong> antibioticsactive against gram-positive bacteria, but resistant strains<br />

have been reported. Lm is considered universally susceptible <strong>to</strong> antibiotics active<br />

against Gram-positive bacteria except 3 rd generation cephalosporins, but resistant<br />

strains have been reported. We have studied retrospectively the antimicrobial resistance<br />

<strong>of</strong> Lm human strains isolated between 1926, the year <strong>of</strong> its discovery, and<br />

2007. 4 816 human Lm strains were picked randomly in the strain collection <strong>of</strong> the<br />

WHO collaborating for Listeria. Their susceptibility <strong>to</strong> 23 clinically relevant antibiotics<br />

was determined by disc diffusion. MICs <strong>of</strong> fluoroquinolones and penicillins<br />

were determined by E-test. All resistant strains were typed by PFGE and<br />

those displaying a unique pr<strong>of</strong>ile were screened for the presence <strong>of</strong> antibiotic resistance<br />

genes by PCR. The mechanisms leading <strong>to</strong> detected resistance was studied<br />

for each resistant strain. Sixty-one human strains (1.27 %) were resistant <strong>to</strong> one<br />

or more antibiotics. Resistance <strong>to</strong> the tetracyclines (n = 34) and fluoroquinolones<br />

(n = 20) was more common and emerged since the late 80s and the late 90s, respectively.<br />

Resistance <strong>to</strong> fluoroquinolones was attributed <strong>to</strong> active efflux. Resistance<br />

<strong>to</strong> the tetracyclines was encoded by tet(M), presumably carried by a conjugative<br />

transposon in the 14 strains which also harboured the int gene. We also<br />

detected the first human isolate with high-level resistance <strong>to</strong> trimethroprim, due<br />

<strong>to</strong> the dfrD gene. Although no penicillin-resistant strain was identified, MICs have<br />

significantly increased since 1926 with values up <strong>to</strong> 2 µg/mL. In conclusion, L.<br />

monocy<strong>to</strong>genes has not acquired significant clinically-relevant antibiotic resistance.<br />

Importantly, no resistance <strong>to</strong> amoxicillin and gentamicin, the recommended<br />

antibiotic regimen for listeriosis, was found. Yet, significant increase <strong>of</strong> penicillin<br />

MICs and the description <strong>of</strong> a strain resistant <strong>to</strong> trimethoprim reinforce the need<br />

for surveillance and are in favour <strong>of</strong> systematic antibiotic susceptibility testing including<br />

penicillin MICs determination.<br />

Today and <strong>to</strong>morrow: The molecular epidemiology <strong>of</strong> listeriosis in the UK<br />

Grant, K., Amar, C., Ma<strong>to</strong>s, J., Mook, P., Little, C. and Gillespie, I.<br />

Health Protection Agency, UK<br />

Listeriosis is a rare but severe foodborne infection caused by Listeria monocy<strong>to</strong>genes:<br />

in the UK it is the leading cause <strong>of</strong> death due <strong>to</strong> a foodborne pathogen. Since<br />

2000 t<strong>here</strong> has been an increase in the number <strong>of</strong> cases associated with patients<br />

aged > 60 years presenting with bacteraemic infection in the UK and in other European<br />

countries and, yet, the reason for this increase remains <strong>to</strong> be established. In<br />

the UK, the number <strong>of</strong> pregnancy related cases <strong>of</strong> listeriosis has remained relatively<br />

stable over the past few years, although, more recently t<strong>here</strong> has been a noticeable<br />

increase in the proportion <strong>of</strong> cases describing themselves as <strong>of</strong> ethnic origin.<br />

Particularly concerning is that since the beginning <strong>of</strong> 2009, t<strong>here</strong> has been an<br />

increase in the <strong>to</strong>tal number <strong>of</strong> pregnancy related cases. Because the incubation<br />

period for listeriosis is long and those at risk <strong>of</strong> infection are <strong>of</strong>ten ill with serious<br />

underlying conditions the investigation <strong>of</strong> individual cases and outbreaks is <strong>of</strong>ten<br />

difficult with an increased emphasis on microbiological evidence. Approximately<br />

170 human and 900 food Listeria monocy<strong>to</strong>genes isolates from across England and<br />

Wales are received annually by the Labora<strong>to</strong>ry <strong>of</strong> Gastrointestinal Pathogens. All<br />

isolates are identified and characterised using molecular methods including PCR–<br />

based serotyping and amplified fragment length polymorphism. This data is used<br />

in conjunction with standardised epidemiological data <strong>to</strong> confirm and link sporadic<br />

cases; <strong>to</strong> identify and investigate clusters and outbreaks; <strong>to</strong> provide information<br />

on routine food testing as well as for national surveillance purposes such as<br />

identifying risk exposures for human infection. This paper will demonstrate, with<br />

recent examples from the UK, the contribution <strong>of</strong> molecular typing <strong>to</strong> epidemiological<br />

investigations and will outline a new direction for typing Listeria monocy<strong>to</strong>genes<br />

which will not only be discrimina<strong>to</strong>ry but provide additional information<br />

on the relatedness <strong>of</strong> strains and their pathogenic potential.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Use <strong>of</strong> Fluorescent Amplified Fragment Length Polymorphism<br />

for improved typing <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Ma<strong>to</strong>s, J., Amar, C., Desai, M., Cross, L. and Grant, K.<br />

Health Protection Agency, UK<br />

Listeria monocy<strong>to</strong>genes is a ubiqui<strong>to</strong>us, psychrotrophic, Gram positive foodborne<br />

pathogen which has the ability <strong>to</strong> persist in food processing environments and<br />

grow in a range <strong>of</strong> ready-<strong>to</strong>-eat foods. It causes the disease listeriosis which mainly<br />

affects those with a weakened immune system including pregnant women,<br />

neonates, the elderly and immunocompromised individuals. Although a rare disease<br />

it has a high mortality rate (20 – 30 %) and is currently responsible for more<br />

deaths than any other foodborne pathogen in the UK. Typing <strong>of</strong> L. monocy<strong>to</strong>genes<br />

strains is critical <strong>to</strong> investigating both sporadic cases and outbreaks enabling the<br />

identification <strong>of</strong> the food vehicle and tracing the origin <strong>of</strong> contamination. The success<br />

<strong>of</strong> any typing method depends on its discrimina<strong>to</strong>ry ability, reproducibility,<br />

cost <strong>of</strong> the procedure and speed <strong>to</strong> getting results. At present Pulse field Gel Electrophoresis<br />

is generally accepted as the gold standard typing method for epidemiological<br />

investigations involving L. monocy<strong>to</strong>genes. However, it is not without its<br />

drawbacks including being lengthy and labour intensive <strong>to</strong> perform, making it unsuitable<br />

for real time analysis. PFGE also suffers from a lack <strong>of</strong> interlabora<strong>to</strong>ry<br />

reproducibility and the analysis <strong>of</strong> pr<strong>of</strong>iles is open <strong>to</strong> subjective interpretation.<br />

This study investigated the use <strong>of</strong> Fluorescent Amplified Fragment Length Polymorphism<br />

(FluAFLP) for typing 842 isolates <strong>of</strong> L. monocy<strong>to</strong>genes from clinical<br />

cases, foods and food processing environments using two restriction enzymes<br />

(HindIII and HhaI). All isolates were previously subtyped by multiplex PCR-based<br />

serotyping and a 133 subset had also been previously typed by PFGE using AscI.<br />

fAFLP was found <strong>to</strong> be highly discrimina<strong>to</strong>ry (D.I. = 0.987), rapid <strong>to</strong> perform, reproducible<br />

and readily amenable <strong>to</strong> au<strong>to</strong>mation. Cluster analysis <strong>of</strong> isolates<br />

showed that fAFLP efficiently separated lineage I isolates from those <strong>of</strong> lineage<br />

II and that t<strong>here</strong> was a high level <strong>of</strong> interlineage discrimination.<br />

Towards an application for field samples <strong>of</strong> a multipathogen platform<br />

for molecular detection <strong>of</strong> raw milk pathogens<br />

Omiccioli, E1, Amagliani, G. 2, Brandi, G. 2, Tonucci, F. 3, Foglini, M. 3 and Magnani, M. 2<br />

1. Diatheva s.r.l., Italy<br />

2. Department <strong>of</strong> Biomolecular Science, University <strong>of</strong> Urbino, Italy<br />

3. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale Umbria e Marche, Italy<br />

Listeria monocy<strong>to</strong>genes, <strong>to</strong>gether with Salmonella spp. and Escherichia coli O157,<br />

are pathogens associated with various food categories, included raw milk, which<br />

commercialisation for human direct use has been encompassed by the<br />

853/2004/EC. Since this product can constitute a risk for human health, the absence<br />

<strong>of</strong> such bacteria should be assessed by sensitive and specific diagnosis. Molecular<br />

diagnostic tests in multiplex format proved <strong>to</strong> be useful for the rapid identification<br />

<strong>of</strong> food contaminations caused by more than one microbial species. As<br />

part <strong>of</strong> a European Research project (DIAGNOSIS), we developed an enrichment<br />

multiplatform Real-Time PCR, including an internal amplification control ensuring<br />

the reliability <strong>of</strong> results, for the detection <strong>of</strong> all three pathogens. The assay<br />

combines an enrichment step in a medium properly formulated for the simultaneous<br />

growth <strong>of</strong> target pathogens (Multipathogen enrichment medium), a DNA<br />

isolation method, and a multiplex Real-Time PCR detection system based either on<br />

dual-labelled probes (MultipathogenFLUO), or on melting curve analysis (MultipathogenHRM).<br />

The entire system enables the detection <strong>of</strong> as few as 1 CFU <strong>of</strong> each<br />

pathogen in 125 ml milk in only two working days and it is now available from<br />

Diatheva srl, Fano, Italy. An application study for field samples is currently in<br />

progress. A local dairy farm, licensed <strong>to</strong> the direct selling <strong>of</strong> raw milk through au<strong>to</strong>matic<br />

distribu<strong>to</strong>rs, was selected for the collection <strong>of</strong> raw milk samples and swabs<br />

from filters located between milking rooms and s<strong>to</strong>rage tanks. Samples are<br />

analysed by the <strong>of</strong>ficial reference ISO methods and also culture-enriched in the<br />

Multipathogen enrichment medium. Aliquots from all enriched cultures are subjected<br />

<strong>to</strong> DNA column-based extraction and PCR amplified by both the MultipathogenHRM<br />

and MultipathogenFLUO kits. The comparison <strong>of</strong> results will allow<br />

the estimation <strong>of</strong> method equivalence and the compliance <strong>of</strong> the molecular kits<br />

with law provisions, contributing <strong>to</strong> promote the adoption <strong>of</strong> PCR-based methods<br />

alongside the traditional diagnostic procedures.<br />

R E F E R E N C E<br />

C / P<br />

92<br />

R E F E R E N C E<br />

C / P<br />

93


R E F E R E N C E<br />

C / P<br />

94<br />

R E F E R E N C E<br />

C / P<br />

95<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Detection and recovery <strong>of</strong> Listeria species from stainless steel<br />

Boone, R., Iugovaz, I, Trottier, Y.-L. and Pagot<strong>to</strong>, F.<br />

Health Canada<br />

The 2008 Canadian listeriosis outbreak in ready <strong>to</strong> eat meat raised many issues<br />

over the safety and reliability <strong>of</strong> the national food supply, including the reliability<br />

<strong>of</strong> detection methodologies aimed at food-contact surfaces made <strong>of</strong> stainless steel.<br />

Five methods pertaining <strong>to</strong> the recovery <strong>of</strong> Listeria species including L. monocy<strong>to</strong>genes<br />

from stainless steel were assessed. A cocktail <strong>of</strong> L. monocy<strong>to</strong>genes (serovars<br />

1/2a, 1/2b, 4b), L. innocua, plus a Gram-positive organism (Staphylococcus aureus)<br />

was inoculated at different levels on<strong>to</strong> stainless steel. Inocula were recovered using<br />

MFLP-41 and assessed using different detection methods, including Petrifilm<br />

(3M), BAX Genus Listeria (Qualicon), BAX Listeria monocy<strong>to</strong>genes, and four<br />

VIDAS kits (BioMérieux) (LDUO, LIS, LMO2 and LSX). Direct plating on<strong>to</strong> selective<br />

and chromogenic agars were done simultaneously from each broth and each<br />

kit was compared <strong>to</strong> the “gold standard”, MFHPB-30. Experiments using skim<br />

milk powder assessing cell viability revealed a consistent 1-log die-<strong>of</strong>f after 24 h.<br />

The VIDAS strips DLIS, LIS and LSX had 74 positive samples out <strong>of</strong> 84 tested following<br />

52 hours <strong>of</strong> incubation in LX Broth while the BAX Genus Listeria had 72<br />

positive samples out <strong>of</strong> the 82 tested. The VIDAS strips DLMO and LMO2 yielded<br />

52 and 51 positive samples out <strong>of</strong> 79 tested, respectively; w<strong>here</strong>as the BAX L. monocy<strong>to</strong>genes<br />

detected 50 <strong>of</strong> 81 samples tested. Petrifilm modifications <strong>to</strong> increase incubation<br />

time from 30 h <strong>to</strong> 48 h are being suggested due <strong>to</strong> 27 out <strong>of</strong> 84 samples<br />

were negative at 30 h but positive at 48. Differences amongst methods were noted<br />

at the single cell inoculum level. None <strong>of</strong> the methods had false positive results.<br />

Through the study <strong>of</strong> different rapid detection methods, options can be brought<br />

forward <strong>to</strong> improve and ensure optimal conditions for Listeria detection (and recovery)<br />

from stainless steel environments.<br />

Natural carriage <strong>of</strong> Listeria in fresh-water fish in Russia<br />

Egorova, I. 1, Voronin, M. 2, Selyaninov, Y. 1 and Kolbasov, D. 1<br />

1. National Institute <strong>of</strong> Veterinary Virology and Microbiology Pokrov Russia<br />

2. National Park “Zavidovo”, Russia<br />

To study the extent <strong>of</strong> Listeria natural carrier stage in fresh-water fish, the<br />

Ivankovo water s<strong>to</strong>rage reservoir was observed. In the period <strong>of</strong> 2006 <strong>to</strong> 2008, 642<br />

fish specimens belonging <strong>to</strong> 18 fresh-water fish species were examined. In <strong>to</strong>tal, 34<br />

listerial cultures <strong>of</strong> monocy<strong>to</strong>genes and innocua species were isolated in fish, 35,3 %<br />

<strong>of</strong> them belong <strong>to</strong> L. monocy<strong>to</strong>genes. Listeria isolation rate was 5,38 %. L. monocy<strong>to</strong>gens<br />

were isolated predominantly from plant-feeder fish like bream, crucian<br />

carp and silver bream. L. innocua - in bream, pike, crucian carp, redeye and perch<br />

<strong>of</strong> all age groups. Listeria carriers were found most <strong>of</strong>ten in areas w<strong>here</strong> animal<br />

farm buildings and treatment facilities were located. The primary Listeria locations<br />

in fish were muscle tissues, gill rakers and parenchyma<strong>to</strong>us organs. In two<br />

cases L. innocua were isolated from fresh-water fish intestines. The peak <strong>of</strong> Listeria<br />

carrying in fish falls on summer periods (40 %). In autumn and winter periods<br />

the detection rates were 28 %. In spring the index dropped <strong>to</strong> 4 %. All the cultures<br />

are quite typical specimens <strong>of</strong> their species. L. monocy<strong>to</strong>genes cultures produced<br />

hemolysins and phospholipases, were agglutinated with specific sera and lysed<br />

with phage L2A, and induced purulent conjunctivitis in guinea pigs. L. innocua<br />

cultures were hemo- and phospholipase-negative and were agglutinated with a<br />

serum <strong>of</strong> serogroup II, were lysed with phages L4A and did not induce mice death<br />

due <strong>to</strong> intraperi<strong>to</strong>neal inoculation at a dose <strong>of</strong> 10 9 m.b. The experimentally determined<br />

diversity <strong>of</strong> the <strong>of</strong> Sma I pulse electrotypes suggests t<strong>here</strong> are both associated<br />

and independent Listeria reservoirs and t<strong>here</strong> is a circulation <strong>of</strong> several clone<br />

variants <strong>of</strong> the pathogen in the fresh-water fish population.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Prevalence <strong>of</strong> Listeria in wild fauna in Russia<br />

Egorova, I. 1, Fertikov, V. 2 and Kolbasov, D. 1<br />

1. National Institute <strong>of</strong> Veterinary Virology and Microbiology Pokrov Russia<br />

2. National Park “Zavidovo”, Russia<br />

Moni<strong>to</strong>ring <strong>of</strong> Listeria in environmental objects and also among ungulate wildlife<br />

inhabiting forest hunting ranges <strong>of</strong> the Central region <strong>of</strong> Russia has been carried<br />

out. 662 samples <strong>of</strong> soil from feeding grounds, decaying plant residues, forages<br />

and wildlife faeces, and also brain, muscle tissue and/or parenchymal organs <strong>of</strong><br />

the axis deer, the maral, the elk and the wild boar, were tested and 68 listerial cultures<br />

were found. The organism carriage indices for the axis deer, the maral and<br />

the wild boar in the areas <strong>of</strong> Tver, Moscow and Kaluga regions were at a similar<br />

level making up 2.0 <strong>to</strong> 2.7 %. In Vladimir region the indices turned out <strong>to</strong> be a little<br />

higher gaining 12 % for the wild boar and 5.45 % for the axis deer, which is most<br />

likely due <strong>to</strong> some differences in the forms <strong>of</strong> economic activities and inadequate<br />

observance <strong>of</strong> sanitary-and-hygienic requirements. Examinations <strong>of</strong> elk faeces revealed<br />

Listeria in none <strong>of</strong> the cases which is probably explained by specificity <strong>of</strong><br />

the elk nutrition. Examinations <strong>of</strong> environmental objects showed that all <strong>of</strong> them<br />

were <strong>to</strong> a variable extent contaminated by Listeria <strong>of</strong> both pathogenic and nonpathogenic<br />

species. Listeria were mainly isolated in soil samples taken from feeding<br />

grounds used for seasonal feeding <strong>of</strong> the wildlife. In samples from wild animals<br />

Listeria were found in two cases in axis deer brain and lymph node, and also<br />

in boar liver and lung. The isolate taken from the axis deer is <strong>of</strong> a certain research<br />

interest, as according <strong>to</strong> all the phenotypic indications it was classified <strong>to</strong> L. innocua,<br />

although investigation <strong>of</strong> its genetic structure showed presence <strong>of</strong> pathogenicity<br />

genes typical for pathogenic Listeria species in its genome. The remaining<br />

isolates <strong>of</strong> Listeria found in the course <strong>of</strong> the moni<strong>to</strong>ring investigations belonged<br />

basically <strong>to</strong> two species, namely L. monocy<strong>to</strong>genes and L. innocua.<br />

Quantitative assessment <strong>of</strong> the exposure <strong>to</strong> Listeria monocy<strong>to</strong>genes<br />

from s<strong>of</strong>t-ripened cheese consumption in North America: a joint<br />

FDA/Health Canada project<br />

Gendel, S. 1, Pouillot, R.1, Murray, C. 1, Farber, J. 2, Ross, W. 2, Couture, H. 2 and Jean, A. 2<br />

1. Food and Drug Administration, USA<br />

2. Health Canada<br />

The United States and Canada continue <strong>to</strong> experience sporadic illnesses and outbreaks<br />

<strong>of</strong> listeriosis associated with the consumption <strong>of</strong> cheese, particularly s<strong>of</strong>t<br />

and s<strong>of</strong>t-ripened cheese. T<strong>here</strong>fore, the US Food and Drug Administration and<br />

Health Canada jointly conducted a risk assessment <strong>to</strong> assess the public health impact<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes in s<strong>of</strong>t-ripened cheese in the US and Canada, and <strong>to</strong><br />

evaluate the impact <strong>of</strong> current and potential mitigation and control strategies. To<br />

support this risk assessment, a probabilistic model was developed <strong>to</strong> assess<br />

changes in L. monocy<strong>to</strong>genes prevalence and levels from “farm <strong>to</strong> fork” in multiple<br />

scenarios. This model evaluated the impact <strong>of</strong> contamination <strong>of</strong> the milk used for<br />

cheese making, in-plant environmental contamination, partitioning, mixing, bacterial<br />

growth and bacterial inactivation though the complete product pathway.<br />

Model calculations and parameters for a base model were derived from data in the<br />

published literature and from specific databases from both countries. The general<br />

framework <strong>of</strong> the model was a second-order simulation that allowed separate evaluation<br />

<strong>of</strong> variability and data uncertainty. An original back-calculation was used <strong>to</strong><br />

evaluate the frequency and level <strong>of</strong> environmental in-plant contamination based<br />

on published data on L. monocy<strong>to</strong>genes prevalence and levels in s<strong>of</strong>t-ripened<br />

cheeses at retail in the US. This back-calculation model was integrated using the<br />

Monte-Carlo method. The results <strong>of</strong> the back calculation suggest that in-plant<br />

contamination is infrequent and generally leads <strong>to</strong> a low levels <strong>of</strong> L monocy<strong>to</strong>genes<br />

per cheese (< 25 cfu in each whole 250 g cheese). Because data on production levels<br />

and practices for the cheese making industry are limited, the model was used <strong>to</strong><br />

compare exposure estimates obtained using the base model <strong>to</strong> estimates obtained<br />

using alternate scenarios. For example, scenario analysis suggested that, for contaminated<br />

cheese, the level <strong>of</strong> exposure is highly sensitive <strong>to</strong> s<strong>to</strong>rage conditions.<br />

R E F E R E N C E<br />

C / P<br />

96<br />

R E F E R E N C E<br />

C / P<br />

97


R E F E R E N C E<br />

C / P<br />

98<br />

R E F E R E N C E<br />

C / P<br />

99<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Human listeriosis in England, 2001-2007: association with neighbourhood<br />

deprivation<br />

Gillespie, I. A., Mook, P., Little, C. L., Grant, K. and McLauchlin, J.<br />

Health Protection Agency, UK<br />

Listeriosis is a rare but severe disease affecting mostly the elderly and the immunocompromised,<br />

and, <strong>to</strong> a lesser extent, pregnant women, unborn or newborn<br />

infants. Despite a high mortality rate, the socioeconomics <strong>of</strong> listeriosis has not<br />

been studied in detail, meaning that health inequalities which might exist in relation<br />

<strong>to</strong> this disease are not apparent. Labora<strong>to</strong>ry surveillance data on cases <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes infection reported in England between 2001 and 2007 were<br />

linked <strong>to</strong> indices <strong>of</strong> deprivation and denomina<strong>to</strong>r data using patients’ postcode.<br />

Incidence relative <strong>to</strong> increasing quintiles <strong>of</strong> deprivation was calculated by fitting<br />

generalized linear models whilst controlling for population. Patient exposure data<br />

were scrutinised and compared with commercial food purchasing denomina<strong>to</strong>r<br />

data <strong>to</strong> further quantify the risk. For all patient groups, listeriosis incidence was<br />

highest in the most deprived areas <strong>of</strong> England when compared <strong>to</strong> the most affluent,<br />

and cases were more likely <strong>to</strong> purchase foods from convenience s<strong>to</strong>res or from<br />

local services (bakers, butchers, fishmongers and greengrocers) than the general<br />

population. Patients’ risk pr<strong>of</strong>ile also changed with increasing neighbourhood deprivation.<br />

With increased life expectancy and rising food prices, food poverty could<br />

become an increasingly important driver for foodborne disease in the future.<br />

Whilst Government policy should continue <strong>to</strong> focus on small food businesses <strong>to</strong><br />

ensure sufficient levels <strong>of</strong> food hygiene expertise, tailored and targeted food safety<br />

advice on the avoidance <strong>of</strong> listeriosis is required for all vulnerable groups. Failure<br />

<strong>to</strong> do so may enhance health inequality across socioeconomic groups.<br />

Characterisation <strong>of</strong> antibodies for use in an antibody-based sensor for<br />

on-site detection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

Gilmartin, N., Hearty, S. and O’ Kennedy, R.<br />

School <strong>of</strong> Biotechnology and National Centre for Sensor Research, Dublin, Ireland<br />

L. monocy<strong>to</strong>genes is a facultative anaerobic, non-sporing, Gram-positive, motile,<br />

rod-shaped bacterium and is an important food-borne pathogen. It is a major<br />

causative agent <strong>of</strong> listeriosis, an infection which can kill vulnerable people such<br />

as the elderly, newborns, pregnant women and people suffering from immunocompromising<br />

diseases. The current methods for sampling and detection <strong>of</strong> L.<br />

monocy<strong>to</strong>genes can result in extensive treatment times, limited sensitivity and<br />

very low recovery rates <strong>of</strong> the microorganism (due <strong>to</strong> the formation <strong>of</strong> bi<strong>of</strong>ilms).<br />

Antibody-based sensors permit the rapid and sensitive analysis <strong>of</strong> a range <strong>of</strong><br />

pathogens and associated <strong>to</strong>xins. They <strong>of</strong>fer many advantages such as improved<br />

sensitivity, real-time analysis and quantification. An anti-L. monocy<strong>to</strong>genes monoclonal<br />

antibody (2B3) was isolated from a panel <strong>of</strong> hybridomas produced using L.<br />

monocy<strong>to</strong>genes serotype 1/2a as the immunogen. This antibody reacted with L.<br />

monocy<strong>to</strong>genes serotypes 1a, 1/2a, 1/2b, 1/2c, 3a, 4a and 4b but did not react with related<br />

Listeria spp. and non-Listeria spp. The mAb was further evaluated in both<br />

enzyme-linked-immunosorbent assay (ELISA) and fluorescence-linkedimmunosorbent<br />

assay (FLISA)-based formats and in a surface plasmon resonance<br />

based biosensor platform. This antibody will be exploited for use in an antibodybased<br />

sensor as part <strong>of</strong> the BIOLISME project which aims <strong>to</strong> develop a system <strong>to</strong><br />

moni<strong>to</strong>r the contamination levels <strong>of</strong> L. monocy<strong>to</strong>genes in industrial food-producing<br />

plants. Preliminary work on the use <strong>of</strong> this antibody in the detection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

in bi<strong>of</strong>ilms will also be presented.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Food Investigation <strong>of</strong> sporadic cases <strong>of</strong> neuroinvasive listeriosis<br />

Goulet, V. 1, Leclercq, A. 2, Laurent, E., King, L., Dusch, V., Salem, S., Vaillant, V. 1,<br />

Chenal-Francisque, V. 2, de Valk, H. 1 and Pihier, N. 3<br />

1. Institut de Veille Sanitaire, France<br />

2. National Reference Centre and WHO collaborating centre for Listeria, France<br />

3. Direction générale de l’Alimentation, France<br />

Listeriosis in France is manda<strong>to</strong>ry notified. For each case, a detailed food his<strong>to</strong>ry is<br />

collected. For cases with central nervous system involvement who consent, samples<br />

are taken from foods and the environment in the kitchen <strong>of</strong> their residence or<br />

<strong>of</strong> hospital for patients hospitalized before disease onset. We present results <strong>of</strong><br />

investigations 2003-2008. Samples are taken by <strong>of</strong>ficers <strong>of</strong> inspection services <strong>of</strong><br />

the ministry <strong>of</strong> agriculture, within 14 days after diagnosis <strong>of</strong> the case, and sent <strong>to</strong><br />

their authorised labora<strong>to</strong>ries. Strains <strong>of</strong> Listeria monocy<strong>to</strong>gnes (Lm) isolated from<br />

these samples are sent <strong>to</strong> the National Reference Centre who performs PCR<br />

serotype, and PFGE (Apa1 and Sma1) typing <strong>of</strong> the food and patient isolates. Lm<br />

contaminated food is traced back with further investigations at the production<br />

sites. 140 cases have been investigated. Lm was isolated in 63 (45 %) investigations,<br />

from cheeses, meat products, and the environment. For 23 investigations,<br />

the food or environmental strains had the undistinguishable PFGE pr<strong>of</strong>iles <strong>of</strong> the<br />

patient strain. At 2 occasions, the patient strains were isolated from the kitchen<br />

environment in their hospital. In 3 other investigations, the patient strain was also<br />

isolated at the production site <strong>of</strong> the contaminated food. At one occasion, the early<br />

investigation <strong>of</strong> a case later shown <strong>to</strong> be part <strong>of</strong> a cluster, allowed the prompt investigation<br />

in a cheese producing fac<strong>to</strong>ry and the confirmation <strong>of</strong> the source <strong>of</strong><br />

the outbreak. Food investigation <strong>of</strong> listeriosis cases are <strong>of</strong> interest because they<br />

can lead <strong>to</strong> control measures in contaminated production sites and limit the size <strong>of</strong><br />

outbreaks <strong>of</strong> listeriosis by rapidly identifying their sources. However the yield <strong>of</strong><br />

these investigations is limited because numerous unpacked food products in the<br />

patient’s refrigera<strong>to</strong>rs can not be easily identified. An important finding is that<br />

Lm strains can colonise the environment <strong>of</strong> hospital kitchens.<br />

Evaluation <strong>of</strong> the antimicrobial activity <strong>of</strong> Vaccinium myrtillus, Prunus<br />

domestica and Myrtus communis against Listeria monocy<strong>to</strong>genes<br />

Serio, A., Di Pasquale, F. and Paparella, A.<br />

Department <strong>of</strong> Food Science, University <strong>of</strong> Teramo, Mosciano Stazione (TE), Italy<br />

Essential oils and plant extracts have been increasingly used as natural food<br />

preservatives. They are considered consumer-friendly and can be used without<br />

any specific authorization or labelling change. In this study, three aqueous plant<br />

extracts were evaluated for their antimicrobial activity against 45 Listeria monocy<strong>to</strong>genes<br />

strains. Among blueberry, plum and myrtle extracts, the latter was the<br />

most effective against the strains isolated from smoked salmon and meat products,<br />

with MIC (Minimal Inhibi<strong>to</strong>ry Concentration) values generally ranging from<br />

0.5 <strong>to</strong> 2.0 %. Interesting results were also obtained with the plum extract, although<br />

with MIC values from 1 <strong>to</strong> 4 %. Blueberry was effective only at high concentration<br />

levels, even above 10 %. Isolates from the smoked salmon industry were generally<br />

less sensitive, suggesting that resistance <strong>to</strong> plant extracts may somehow be related<br />

<strong>to</strong> strain origin. A 3-fac<strong>to</strong>r-5-level Central Composite Design was used <strong>to</strong> evaluate<br />

the extracts activity in combination with compounds commonly present in some<br />

food products, w<strong>here</strong> they can constitute a hurdle for bacterial growth. Myrtle extract<br />

activity was boosted by lactic acid and salt, while ATCC7644 strain, used as<br />

control, was particularly susceptible <strong>to</strong> all combinations. These results are in line<br />

with our previous research on volatile compounds, w<strong>here</strong> L. monocy<strong>to</strong>genes was<br />

inhibited at low extract concentrations, without significantly affecting food flavor.<br />

Further studies will be carried out <strong>to</strong> evaluate the potential <strong>of</strong> myrtle extract in<br />

food biopreservation.<br />

R E F E R E N C E<br />

C / P<br />

100<br />

R E F E R E N C E<br />

C / P<br />

101


R E F E R E N C E<br />

C / P<br />

102<br />

R E F E R E N C E<br />

C / P<br />

103<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

High throughput quantitative detection <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in food by RealTime PCR<br />

Cammà, C., Ancora, M., Rizzi, V., Sperandii, A., Prencipe, V. and Migliorati, G.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

Listeria monocy<strong>to</strong>genes is one <strong>of</strong> the most important food-borne pathogens and<br />

can cause a severe disease in human. The Commission Regulation (EC) N°<br />

2073/2005 <strong>of</strong> 15 November 2005 on microbiological criteria for foodstuffs allows<br />

the presence <strong>of</strong> 100 CFU/g in some ready-<strong>to</strong>-eat products. Conventional bacteriological<br />

methods for the detection and quantification <strong>of</strong> L. monocy<strong>to</strong>genes are laborious<br />

and time-consuming. The purpose <strong>of</strong> this study was <strong>to</strong> develop a rapid and<br />

au<strong>to</strong>mated method for detection and quantification <strong>of</strong> L. monocy<strong>to</strong>genes in meat.<br />

The target <strong>of</strong> the RealTime PCR was a portion <strong>of</strong> hly gene and the procedure consists<br />

<strong>of</strong> two au<strong>to</strong>mated steps: DNA extraction, and set up <strong>of</strong> a fast RealTime PCR<br />

mix. The whole process could be completed within few hours minimizing the risk<br />

<strong>of</strong> carry-over contamination. To evaluate the inhibi<strong>to</strong>r effect <strong>of</strong> food matrix on the<br />

PCR an internal amplification control was included in each reaction and coamplified<br />

using the same primer pair used for the target amplification. The specificity <strong>of</strong><br />

the assay was confirmed on 3 L. monocy<strong>to</strong>genes strains and 26 other bacterial<br />

strains. The method was able <strong>to</strong> detect 2 genomic equivalents (GE) in 33/42 replicates<br />

(LOD= 4.8 GE/reaction with a probability <strong>of</strong> 95 %) and the efficiency resulted<br />

between 90 % and 98 % calculated by analyzing 7 standard curves based on six 10fold<br />

dilutions starting from a 10 7 CFU/ml broth culture <strong>of</strong> L. monocy<strong>to</strong>genes. The<br />

quantification range was linear over 6 log units with a square regression coefficient<br />

> 0.99. Application <strong>of</strong> the assay was assessed with 12 artificially contaminated<br />

samples <strong>of</strong> canned meat and with 8 real samples (smoked salmon, fresh fish<br />

and salami). The microbial load was tested in comparison with a pour plate counting<br />

method. The RealTime PCR was able <strong>to</strong> define an accurate quantification up <strong>to</strong><br />

10 3 CFU/g both in naturally and in artificially contaminated samples.<br />

Bibliographic study concerning procedures for preparing environmental<br />

samples for analyses, regarding <strong>to</strong> L. monocy<strong>to</strong>genes<br />

Barre, L., Carpentier, B. and Gnanou Besse, N.<br />

AFSSA, France<br />

Listeria monocy<strong>to</strong>genes (L. monocy<strong>to</strong>genes) is a bacterium that can contaminate<br />

foods and cause a mild non-invasive illness or a severe, sometimes life-threatening,<br />

illness. L. monocy<strong>to</strong>genes is widespread in the environment. It can be readily isolated<br />

from humans, domestic animals, raw agricultural commodities, and food processing<br />

environments. Sanitation controls include effective environmental moni<strong>to</strong>ring<br />

programs designed <strong>to</strong> identify and eliminate L. monocy<strong>to</strong>genes in and on<br />

surfaces and areas in the plant. The conditions are likely <strong>to</strong> be present in food fac<strong>to</strong>ries<br />

that may give rise <strong>to</strong> the development <strong>of</strong> persistent L. monocy<strong>to</strong>genes strains.<br />

This persistence can be observed in spite <strong>of</strong> the well applied procedures <strong>of</strong> hygiene.<br />

This document is a bibliographic study concerning procedures for preparing<br />

environmental samples for analyses, regarding <strong>to</strong> L. monocy<strong>to</strong>genes. Different procedures<br />

for collecting environmental samples are described: several procedures<br />

for sampling methods are compared (cot<strong>to</strong>n swab, sterile sponge, composite-ply<br />

tissues, sonication, direct contact agar…). We have also compared different detection<br />

and enumeration methods. We concluded that swabbing is known not <strong>to</strong> detach<br />

all micro. The removal efficiency depends on the swab material. Swabbing in<br />

moistened condition, as processed <strong>here</strong>, is more efficient than dry swabbing. Use <strong>of</strong><br />

ultrasound is still considered as a good method for detaching microorganisms.<br />

However, ultrasounds can be lethal <strong>to</strong> bacteria, particularly when used at low frequency<br />

and are more detrimental <strong>to</strong> attached bacteria than <strong>to</strong> suspended ones. We<br />

concluded that it is important <strong>to</strong> adapt the type <strong>of</strong> sampling <strong>to</strong>ols and techniques<br />

<strong>to</strong> the type <strong>of</strong> surfaces and sampling locations.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Persistent L. monocy<strong>to</strong>genes isolates from Austrian and Irish dairies<br />

show the same phenotypic and genetic background<br />

Stessl, B.<br />

Veterinary University Vienna, Austria<br />

The food borne pathogen Listeria monocy<strong>to</strong>genes continues <strong>to</strong> be <strong>of</strong> major epidemiological<br />

interest <strong>to</strong> food industry and food sciences. The Zoonosis Report<br />

2004 <strong>of</strong> EFSA (European Food Safety Authority) listed 1267 cases <strong>of</strong> listeriosis in<br />

Europe in 2004, resultant 0.3 cases per 100 000 capita. Human infections with L.<br />

monocy<strong>to</strong>genes are rare, but are important due <strong>to</strong> the high mortality rate. This<br />

study is focused on the PFGE (Pulsed Field Gel Electrophoresis) and FTIR<br />

(Fourier Transform Infrared Spectroscopy) typing <strong>of</strong> L. monocy<strong>to</strong>genes strains isolated<br />

at distant sampling points (1996 – 2007) in the Austrian and Irish dairy chain.<br />

The aim <strong>of</strong> this research was <strong>to</strong> analyse if genetically indistinguishable (“persistent“)<br />

L. monocy<strong>to</strong>genes clones can be differentiated with respect <strong>to</strong> their phenotype<br />

(adaptation <strong>to</strong>wards sanitizers and hygienic measures). A set <strong>of</strong> L. monocy<strong>to</strong>genes<br />

isolates from seven different European dairies was analysed with two<br />

epidemiological techniques: PFGE and FTIR. The PFGE method was performed<br />

according the CDC PulseNet Standardized Pro<strong>to</strong>col for Molecular Subtyping <strong>of</strong> L.<br />

monocy<strong>to</strong>genes. Further data analysis was completed by Fingerprinting II Cluster<br />

Analysis using Dice coefficient and UPGMA logarithm (Biorad, Austria). The FTIR<br />

sample preparation and spectra analysis were performed as described by Oberreuter<br />

et al. For data processing the program OPUS FT-IR (Bruker, Germany) was<br />

used. Cluster analysis was performed using Ward’s algorithm. The PFGE as well as<br />

the FTIR cluster analysis suggests the ability <strong>of</strong> short and long-time persistence <strong>of</strong><br />

L. monocy<strong>to</strong>genes isolates in s<strong>of</strong>t, semi-hard and hard cheese manufacturing. An<br />

explanation could be that the tested persistent L. monocy<strong>to</strong>genes isolates were not<br />

exposed <strong>to</strong> selection due <strong>to</strong> sanitation (colonization <strong>of</strong> ecological niches characterized<br />

by a low hygienic pressure). Additional studies will concentrate on the testing<br />

<strong>of</strong> a L. monocy<strong>to</strong>genes strain subset against their disinfectant susceptibility.<br />

Epidemiological data on listeriosis in Portugal: 2003 – 2008<br />

Almeida, G, Magalhães, R., Hogg, T. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

The invasive form <strong>of</strong> listeriosis although rare is a very serious disease with a high<br />

mortality rate. Young, Old, Pregnant, Immuno-compromised, sometimes referred<br />

as YOPI, are identified as at risk groups. However the infection may also occur in<br />

people with no known predisposing fac<strong>to</strong>rs. In Portugal listeriosis is a not a notifiable<br />

disease and is likely <strong>to</strong> be under-reported. In 2003 an incidence <strong>of</strong> 1.4 cases per<br />

million inhabitants was reported. As in other European countries the number <strong>of</strong><br />

cases has increased in the last years. The aim <strong>of</strong> this study is <strong>to</strong> present epidemiological<br />

on human cases <strong>of</strong> listeriosis occurred in Portugal between 2003 and 2008.<br />

It was possible <strong>to</strong> gather 107 cases: 16 % were maternal/neonatal infections and<br />

84% were non-maternal/neonatal infections. Concerning non-maternal/neonatal<br />

cases the sex ratio (M/F) was 1.73, the mean age <strong>of</strong> the patients was 62 years old<br />

with 46 cases (51 %) in individuals aged 65 or more. The microorganism was mainly<br />

isolated from blood (59 %), from CSF (32 %) and from other specimen (9 %). The incidence<br />

<strong>of</strong> listeriosis in Portugal based on voluntary reporting is lower than the European<br />

notification rate (0.3/100000 inhabitants) and similar <strong>to</strong> countries like Austria,<br />

Es<strong>to</strong>nia, Latvia, Slovakia, Slovenia and Spain. The need <strong>of</strong> an active<br />

surveillance system <strong>of</strong> listeriosis was clearly demonstrated.<br />

R E F E R E N C E<br />

C / P<br />

104<br />

R E F E R E N C E<br />

C / P<br />

105


R E F E R E N C E<br />

C / P<br />

106<br />

R E F E R E N C E<br />

C / P<br />

107<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Diversity among Listeria monocy<strong>to</strong>genes isolated from humans<br />

Kalekar, S. 1*, Rodrigues, J. 1, D’Costa, D. 1, Malik, S. V. S. 2, Kalorey, D. R. 3, Chakraborty, T. 4<br />

and Barbuddhe, S. B. 1<br />

1. ICAR Research Complex for Goa, India<br />

2. Indian Veterinary Research Institute, Índia<br />

3. Nagpur Veterinary College, India<br />

4. Institute <strong>of</strong> Medical Microbiology, Justus Liebig University, Germany<br />

Listeria monocy<strong>to</strong>genes is a food borne pathogen associated with severe diseases in<br />

humans and animals. It is associated with neural, visceral and reproductive clinical<br />

entities, leading <strong>to</strong> septicaemia, abortions, stillbirth, meningitis and meningoencephalitis,<br />

especially in individuals at risk. We present the genotypic analysis<br />

<strong>of</strong> 17 L. monocy<strong>to</strong>genes isolates recovered from humans in India during 2006–2009<br />

by multiplex serotyping PCR allowing serovar predictions, conventional serology<br />

and by PFGE. Multiplex-PCR serotyping assay revealed 88.24 % (15/17) <strong>of</strong> the<br />

strains belonging <strong>to</strong> serovar group 4b, 4d, 4e, 11.76 % (2/17) <strong>to</strong> serovar group 1/2b,<br />

3b. Twelve (70.59 %) L. monocy<strong>to</strong>genes isolates were assigned <strong>to</strong> serotype 4b, 2<br />

(11.76 %) serotype 4d, one (5.88 %) serotype 4e and 2 (11.76 %) <strong>to</strong> serotype 1/2b by<br />

serology. Ten ApaI pulsotypes were recognized among the 17 human isolates.<br />

PFGE analysis allowed discrimination among isolates <strong>of</strong> the same serotype and<br />

among isolates from the same sampling areas or those isolated from different<br />

areas. Thus, PFGE <strong>to</strong>gether with multiplex-PCR serotyping allows rapid discrimination<br />

<strong>of</strong> L. monocy<strong>to</strong>genes strains. In addition, the predominance <strong>of</strong> L. monocy<strong>to</strong>genes<br />

serotype 4b is <strong>of</strong> concern, as this serotype has been most frequently associated<br />

with human listeriosis outbreaks.<br />

* Participation Supported by IUFoST.<br />

Characterization <strong>of</strong> Listeria isolated from seafood<br />

Rodrigues, J.*, Kalekar, S., Bhosle, S. N., Doijad, S. and Barbuddhe, S. B.<br />

ICAR Research Complex for Goa, India<br />

Listeria monocy<strong>to</strong>genes, the causative organism <strong>of</strong> listeriosis, is primarily transmitted<br />

<strong>to</strong> humans through contaminated food. While several reports indicate that<br />

fish and fishery products can be frequently contaminated with L. monocy<strong>to</strong>genes,<br />

no major outbreaks associated with these products have been reported. Since fish<br />

and fishery products may be a vehicle for L. monocy<strong>to</strong>genes, it is important <strong>to</strong> have<br />

information on the incidence <strong>of</strong> this pathogen. The presence <strong>of</strong> L. monocy<strong>to</strong>genes<br />

has been documented in the tropical environment but the incidence in fish and<br />

fish products is low. We examined the prevalence, molecular characteristics and<br />

virulence potential <strong>of</strong> L. monocy<strong>to</strong>genes isolates from fishery products <strong>to</strong> gain further<br />

insights on the public health risk caused by this important food borne<br />

pathogen. A <strong>to</strong>tal <strong>of</strong> 221 raw seafood samples were examined for the presence <strong>of</strong><br />

Listeria species. Of these 37 (16.74 %) samples were positive for Listeria species.<br />

Out <strong>of</strong> these, 4 (1.8 %) isolates were confirmed as L. monocy<strong>to</strong>genes. Other species<br />

detected was L. innocua (33). The hlyA gene was detected in all the L. monocy<strong>to</strong>genes<br />

isolates. All the L. monocy<strong>to</strong>genes exhibited phosphatidyl inosi<strong>to</strong>l specific<br />

phospholipase C activity on ALOA agar. Multiplex PCR based serotyping revealed<br />

three L. monocy<strong>to</strong>genes isolates <strong>to</strong> be 1/2a, 1/2c, 3a, 3c group. One isolate belonged<br />

<strong>to</strong> 1/2b, 3b serogroup. The isolates were grouped in<strong>to</strong> three AscI PFGE pr<strong>of</strong>iles.<br />

Prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in fresh seafood is <strong>of</strong> significance as it may contaminate<br />

and persist in the processing environment.<br />

* Participation Supported by IUFoST.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Characterization <strong>of</strong> Listeria species isolated from milk<br />

D’Costa, D. 1*, Bhosle, S. N. 2, Dhuri, R. B. 3, Kalekar, S. 1, Rodrigues, J. 1, Doijad, S. P. 1<br />

and Barbuddhe, S. B. 1<br />

1. ICAR Research Complex for Goa, India<br />

2. Department <strong>of</strong> Microbiology, Goa University, India<br />

3. Goa State Cooperative Milk Producers’ Union Limited, Curti, Ponda, Goa, India<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen commonly found in food and environment.<br />

This study aimed <strong>to</strong> determine the occurrence <strong>of</strong> L. monocy<strong>to</strong>genes in<br />

raw and market milk. Seven hundred sixty seven raw and market milk samples<br />

were collected at different levels <strong>of</strong> collection and processing. The samples taken<br />

included cow milk collected at udder level, from the milk cans, milk receiving centres,<br />

processing unit and market. Samples were enriched in Listeria enrichment<br />

broth and incubated for 48 h, followed by plating on PALCAM agar. Presumptive L.<br />

monocy<strong>to</strong>genes isolates were purified and confirmed by PCR targeting the hly gene.<br />

Overall, 10.56 % <strong>of</strong> the samples (81 <strong>of</strong> 767) were positive for Listeria species. Out <strong>of</strong><br />

these 37 (4.82 %) were confirmed as L. monocy<strong>to</strong>genes. Other Listeria species isolated<br />

were L. innocua (5.47 %), L. ivanovii (0.13 %) and L. grayi (0.13 %). Maximum<br />

isolates were recovered from samples collected from market followed by samples<br />

at milk processing unit. L. monocy<strong>to</strong>genes were serotyped using multiplex PCR<br />

method. A larger proportion <strong>of</strong> isolates (26) belonged <strong>to</strong> group corresponding <strong>to</strong><br />

serovars 1/2a, 1/2c, 3a, and 3c. Serogroup corresponding <strong>to</strong> serovars 4b, 4d and 4e<br />

was detected in two strains while serogroup 1/2b, 3b, 4b, 4d, and 4e was detected in<br />

nine strains. . This study demonstrates the prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in the<br />

raw and market milk and the need for good hygiene practices <strong>to</strong> prevent its entry<br />

in<strong>to</strong> the food chain.<br />

* Participation Supported by IUFoST.<br />

Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in chicken production chain<br />

in Thailand<br />

Kanarat, S., Nijthavorn, N. and Sukhapesna, J.<br />

Veterinary Public Helath Labora<strong>to</strong>ry, Bureau <strong>of</strong> Quality Control <strong>of</strong> Lives<strong>to</strong>ck Products, Department<br />

<strong>of</strong> Lives<strong>to</strong>ck Development Phaya-thai Rd., Bangkok, Thailand<br />

This study was conducted from 2004 <strong>to</strong> 2009 <strong>to</strong> investigate the prevalence <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes in the chicken production chain in Thailand, i.e., from primary<br />

production stage <strong>to</strong> further processing plants. Samples were taken from 43<br />

breeder farms, 32 hatcheries, 1331 broiler farms, 22 slaughterhouses and 22 further<br />

processing plants. Various types <strong>of</strong> samples were collected: fecal drops on the litter,<br />

drinking water, and chicken feed from breeder and broiler farms; swabs and<br />

mecomium from hatcheries; cloacal swabs and fresh chicken meat from slaughterhouses;<br />

and ready-<strong>to</strong>-eat chicken products from further processing plants. The<br />

study showed that t<strong>here</strong> was no L. monocy<strong>to</strong>genes contamination in the chicken<br />

production chain except in slaughterhouses and further processing plants. This<br />

implies that the chickens did not carry the organism in<strong>to</strong> slaughterhouses, and<br />

consequently primary production practices were not responsible for the the contamination<br />

<strong>of</strong> end products. This suggested that the observed L. monocy<strong>to</strong>genes<br />

contamination in 2.5 % and 0.2 % <strong>of</strong> samples <strong>of</strong> fresh chicken meat and ready-<strong>to</strong>eat<br />

chicken products, respectively, was due <strong>to</strong> the breakdowns in the application <strong>of</strong><br />

good hygienic and/or good manufacturing practices.<br />

R E F E R E N C E<br />

C / P<br />

108<br />

R E F E R E N C E<br />

C / P<br />

109


R E F E R E N C E<br />

C / P<br />

110<br />

R E F E R E N C E<br />

C / P<br />

111<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Listeria monocy<strong>to</strong>genes in Ireland – Epidemiology and Molecular Typing<br />

Cormican, M. G. 1, DeLappe, N. 2, McKeown, P. 3 and Garvey, P. 3<br />

1. NUIGalway, Ireland<br />

2. National Salmonella Reference Labora<strong>to</strong>ry Ireland, Ireland<br />

3. Health Protection Surveillance Centre, Ireland<br />

Since January 2004 listeriosis is a notifiable disease in Ireland. Clinicians and labora<strong>to</strong>ry<br />

direc<strong>to</strong>rs must inform the Medical Officer <strong>of</strong> Health <strong>of</strong> all cases diagnosed.<br />

Information is collated by the national Health Protection Surveillance Centre.<br />

Isolates <strong>of</strong> Listeria spp. from human cases and from food products may be voluntarily<br />

submitted <strong>to</strong> the National Salmonella Reference Labora<strong>to</strong>ry. For the years<br />

2004 <strong>to</strong> 2007 11, 12, 7 and 21 (<strong>to</strong>tal =51) cases <strong>of</strong> listeriosis were reported <strong>to</strong> the<br />

HPSC. T<strong>here</strong> were 37 cases categorized as adult or juvenile, 10 pregnancy-related<br />

and 4 neonatal. Cases were distributed through out Ireland with a slight seasonal<br />

peak in Summer in most years. For the years 2004 <strong>to</strong> 2008 the NSRL has received<br />

4, 4, 1, 12 and 14 (<strong>to</strong>tal <strong>of</strong> 35) L. monocy<strong>to</strong>genes isolates from humans and 6, 34, 16,<br />

18 and 12 (<strong>to</strong>tal <strong>of</strong> 86) from foods. Isolates from human cases are predominantly<br />

serotype 4b (n = 24). Isolates from food are predominantly serotype 1/2 (n= 74).<br />

Pulsed field gel electrophoresis (PFGE) performed by the Pulse Net method with<br />

ApaI and AscI enzymes and analysed with Bionumerics s<strong>of</strong>tware ahs allowed the<br />

definition <strong>of</strong> 10 clusters <strong>of</strong> closely related isolates (2 <strong>to</strong> 4 isolates per cluster). The<br />

clusters link isolates from apparently sporadic human cases, including cases occurring<br />

in different years and link isolates from food items with human cases. As<br />

the clusters are small and it is not possible <strong>to</strong> perform typing in real-time epidemiological<br />

links between cases have not been identified. The number <strong>of</strong> cases <strong>of</strong><br />

listeriosis in Ireland has increased in recent years. T<strong>here</strong> is labora<strong>to</strong>ry evidence<br />

linking individual cases and linking cases with food. Real time molecular typing <strong>of</strong><br />

isolates is essential <strong>to</strong> identify linked cases in time frame that would allow timely<br />

epidemiological investigation and intervention.<br />

Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated from human cases<br />

<strong>of</strong> listeriosis occurred in Portugal in 2008<br />

Magalhães, R.*, Barbosa, J., San<strong>to</strong>s, I., Almeida, G. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive foodborne pathogen that can cause serious<br />

infections in humans. High-risk groups include the elderly, immunocompromised<br />

individuals, pregnant women and their neonates Clinical manifestations<br />

<strong>of</strong> invasive human listeriosis include meningitis, encephalitis, late-term spontaneous<br />

abortion, and septicaemia. The fatality rate is high: 20 <strong>to</strong> 30 %. Foodborne<br />

transmission <strong>of</strong> L. monocy<strong>to</strong>genes can also cause a self-limiting acute febrile gastroenteritis<br />

in immunocompetent persons. Listeriosis is an infection <strong>of</strong> considerable<br />

public health concern. In 2008, 27 isolates <strong>of</strong> Listeria monocy<strong>to</strong>genes were recovered<br />

from Portuguese human cases <strong>of</strong> listeriosis and have been characterized<br />

by biotyping (cadmium and arsenic sensitivity), genoserotyping and typing by<br />

pulsed field gel electrophoresis (PFGE) using the enzymes AscI and ApaI. Strains<br />

were aggregated in three multiplex PCR serogroups: IVb (66.7 %), IIb (25.9 %) and<br />

IIa (7.4 %). Concerning biotyping four groups were found: sensitive <strong>to</strong> arsenic and<br />

cadmium (51.9 %), arsenic sensitive and cadmium resistant (14.8 %), resistant <strong>to</strong> arsenic<br />

and sensitive <strong>to</strong> cadmium (29.6 %) and resistant <strong>to</strong> both heavy metals (3.7 %).<br />

Twenty two pulsotypes were indentified. The more prevalent pulsotype aggregates<br />

three isolates, <strong>of</strong> these, two were time and geographical related. Two other pulsotypes<br />

aggregate two time and geographical related isolates. It was shown that seven<br />

pulsotypes (32 %) were also found in previous years.<br />

* Participation Supported by IUFoST.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Post-processing environmental contamination <strong>of</strong> surface-ripened<br />

s<strong>of</strong>t cheese during affinage<br />

D’Amico, D. and Donnelly, C.<br />

University <strong>of</strong> Vermont, USA<br />

Environmental sampling conducted in a cheese aging room identified extensive<br />

contamination with Listeria monocy<strong>to</strong>genes. Further investigation revealed cross<br />

contamination <strong>of</strong> washed rind cheeses aging on various racks. To determine the<br />

extent <strong>of</strong> post-processing contamination <strong>of</strong> cheeses in this facility, two commercially<br />

available PCR assays were investigated as <strong>to</strong>ols for detection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

in naturally contaminated surface-ripened s<strong>of</strong>t cheeses. Methods were<br />

culture confirmed on CHROMagar Listeria. For analysis, cheeses were composited<br />

and 10g sub-samples were assigned <strong>to</strong> an enrichment broth or Butterfield’s<br />

Phosphate buffer for enumeration. Cheeses identified as positive for L. monocy<strong>to</strong>genes<br />

by PCR and/or culture were subsequently enumerated. To achieve a detection<br />

limit <strong>of</strong> ≥ 5 cfu/g, composited samples were diluted 1:5 followed by direct plating<br />

<strong>of</strong> 1 ml on CHROMagar Listeria. Contamination levels ranged from < 5 <strong>to</strong><br />

1,600,000 cfu/g and varied by cheese type with the highest contamination levels<br />

(mean ~356,000 cfu/g) observed in the smallest cheese variety (~200 g) and lower<br />

levels (mean ~1,500-2,500 cfu/g) in the larger cheese types (~400-2000 g). Preliminary<br />

ribotyping <strong>of</strong> random isolates recovered from each <strong>of</strong> the cheese varieties<br />

identified a single lineage II ribotype, DUP-1030B. Based on PCR and cultural<br />

results, 10g samples <strong>of</strong> cheese contaminated at levels as low as < 5 cfu/g were<br />

identified following enrichments <strong>of</strong> 24 or 26 hours or after secondary enrichment<br />

for a <strong>to</strong>tal <strong>of</strong> 44-48 hours. Single enrichment procedures <strong>of</strong> 24 and 26 hours in<br />

Buffered Listeria Enrichment Broth and Oxoid 24LEB, respectively, followed by<br />

plating 100µl on CHROMagar Listeria identified approximately 90 % <strong>of</strong> the true<br />

positive samples. Secondary enrichment in MOPS-BLEB increased the sensitivity<br />

<strong>to</strong> nearly 97 %. PCR detection following single primary enrichment was less sensitive<br />

when compared <strong>to</strong> cultural techniques w<strong>here</strong>as dual enrichment PCR methods<br />

were comparable <strong>to</strong> their cultural counterparts.<br />

Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes isolated<br />

from Portuguese cheeses<br />

Almeida, G., Magalhães, R., San<strong>to</strong>s, I., Barbosa, J., Hogg, T. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

The invasive form <strong>of</strong> listeriosis is a rare but serious disease with a high mortality<br />

rate. Cheese has been associated with several outbreaks <strong>of</strong> human listeriosis involving<br />

large numbers <strong>of</strong> consumers, with a wide economic and public health impact.<br />

To reduce the incidence <strong>of</strong> listeriosis it would be necessary <strong>to</strong> reduce the<br />

contamination level at production and at retail, or <strong>to</strong> inhibit the growth <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes in high-risk foods. Typing <strong>of</strong> bacterial isolates reveals the relationship<br />

among isolates. They can be used <strong>to</strong> determine the primary sources <strong>of</strong> bacterial<br />

contamination and <strong>to</strong> link people who are ill following the consumption <strong>of</strong><br />

contaminated foods <strong>to</strong> the sources <strong>of</strong> bacterial contamination. Eighty isolates <strong>of</strong> L.<br />

monocy<strong>to</strong>genes recovered from cheeses (40 from cheeses purchased at retail, 23<br />

from cheeses ready <strong>to</strong> go <strong>to</strong> the market, 16 from cheeses analysed at Au<strong>to</strong>ridade de<br />

Segurança Alimenter e Económica labora<strong>to</strong>ry and one isolate from cheese in quality<br />

control routine activities) were typed by multiplex PCR serogrouping and by<br />

PFGE using AscI and ApaI restriction enzymes.<br />

Isolates were distributed by four MPCR groups: IIa (11 %), IIb (30 %), IIc (11 %)<br />

and IVb (48 %). Strains involved in human episodes in Portugal were mainly IVb<br />

(72 %) followed by IIb (18 %) and IIa (10 %). The combination <strong>of</strong> the pr<strong>of</strong>iles obtained<br />

with both restriction enzymes (AscI and ApaI) yielded a <strong>to</strong>tal <strong>of</strong> 26 pulsotypes.<br />

Eight <strong>of</strong> these pulsotypes were previously isolated from clinical cases <strong>of</strong> listeriosis<br />

in Portugal. This study highlights the possibility <strong>of</strong> cheeses being<br />

responsible for human cases <strong>of</strong> listeriosis in the country.<br />

R E F E R E N C E<br />

C / P<br />

112<br />

R E F E R E N C E<br />

C / P<br />

113


R E F E R E N C E<br />

C / P<br />

114<br />

R E F E R E N C E<br />

C / P<br />

115<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Prevalence <strong>of</strong> Listeria spp. in retail raw ground beef in Izmir, Turkey:<br />

A comparison <strong>of</strong> standard cultural method and Fluorescent<br />

in situ Hybridization (FISH) technique for detection<br />

Handan Baysal, A.*<br />

Izmir Institute <strong>of</strong> Technology, Department <strong>of</strong> Food Engineering, Izmir, Turkey<br />

Fluorescent in situ hybridization (FISH) is based on the principle that hybridization<br />

<strong>of</strong> a nucleic acid sequence target <strong>of</strong> a microorganism with a specific DNA<br />

probe labeled with a fluorochrome and imaging by a fluorescence microscope.<br />

FISH method using a fluorescent labelled oligonucleotide probe designed from<br />

specific DNA and RNA sequences has proven <strong>to</strong> be a useful <strong>to</strong>ol for the detection <strong>of</strong><br />

a specific microorganism in environmental and clinical samples however the application<br />

<strong>of</strong> the method in food microbiology is investigated by the research studies.<br />

It usually involves four steps: sample fixation and permeabilization; hybridization<br />

<strong>of</strong> the target sequence and the fluorescent probe; stringency washes;<br />

and detection <strong>of</strong> the hybridized cells. In this study the possibility <strong>to</strong> use FISH<br />

method for detecting a food-borne pathogen Listeria spp. that used as food safety<br />

index in ground meat in market place <strong>of</strong> Izmir was evaluated. The work consists <strong>of</strong><br />

two steps. In the first step specified pathogen bacterium was inoculated in ground<br />

beef and the applicability <strong>of</strong> FISH method was evaluated and in the second step the<br />

pathogen was detected in fresh ground beef by standard (conventional) cultural<br />

methods and FISH method. High correlation (r = 0.96) was found between FISH<br />

and the conventional plate count method. The FISH pro<strong>to</strong>col revealed <strong>to</strong> be specific<br />

for Listeria spp. detection, since all bacteria showed a positive hybridization<br />

signal. Results obtained from the direct application <strong>of</strong> the FISH technique <strong>to</strong> cultures<br />

show that it allows the detection <strong>of</strong> Listeria spp. in hours.<br />

* Participation Supported by IUFoST.<br />

Using ListexP100 for Listeria monocy<strong>to</strong>genes detection in foods<br />

Flores Lopes, J. 1, Ferreira Leite, I. 1, Azeredo, J. 2, Gibbs, P. 1 and Teixeira, P. 1<br />

1. CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

2. IBB / Institute for Biotechnology and Bioengineering, Centre <strong>of</strong> Biological, Engineering,<br />

Universidade do Minho, Campus de Gualtar, Braga, Portugal<br />

Food types susceptible <strong>of</strong> being contaminated with Listeria monocy<strong>to</strong>genes, such as<br />

cheese, smoked fish, salads, raw vegetables, sausages and, in general, ready-<strong>to</strong>-eat<br />

foods, pose a serious threat <strong>to</strong> consumers as this pathogen can cause severe diseases.<br />

L. monocy<strong>to</strong>genes can multiply even at refrigeration temperatures, with the minimum<br />

infective dose still remaining unknown. Development <strong>of</strong> efficient and rapid methods<br />

for the a priori detection <strong>of</strong> contamination by this microorganism in foods is <strong>of</strong> great<br />

significance and is required in order <strong>to</strong> ensure public health concerning consumption<br />

<strong>of</strong> foods that are considered <strong>to</strong> be <strong>of</strong> higher contamination risk. The vast majority<br />

<strong>of</strong> Listeria-specific detection methods, developed as alternatives <strong>to</strong> the standard<br />

selective plating methods, possess drawbacks such as a low specificity or efficiency<br />

(which is the case <strong>of</strong> antibody-based assays), or also detect dead cells (in the case <strong>of</strong> assays<br />

that detect genetic material, e.g. PCR, restriction-enzyme pattern-based assays).<br />

The main goal <strong>of</strong> the research work described <strong>here</strong>in was <strong>to</strong> investigate the possibility<br />

<strong>of</strong> using ListexP100 (a commercial bacteriophage-based preparation for the control<br />

<strong>of</strong> Listeria contamination in foods, gently supplied by EBI Food Safety B.V.) for the<br />

design <strong>of</strong> a fast-response detection method for L. monocy<strong>to</strong>genes in foods. The procedure<br />

developed during our research effort encompassed a bacteriophage amplification<br />

assay based on the specificity and lytic activity <strong>of</strong> the broad-host range P100<br />

listeriaphage with plaque formation as the assay end point. Aliquots <strong>of</strong> phage P100<br />

were added <strong>to</strong> the test samples allowing all Listeria cells <strong>to</strong> be infected. After 15 minutes<br />

<strong>of</strong> incubation at 30ºC, <strong>to</strong> allow phage adsorption and eclipse phase with the nucleic<br />

acid delivered in<strong>to</strong> the host cell, a virucidal agent was added in order <strong>to</strong> destroy<br />

all those phages that have not effectively infected a bacterial cell. Among several effective<br />

virucidal agents we selected 4.8 mM FeSO4 and 13 ppm <strong>of</strong> tannic acid and left<br />

standing for 5 min at room temperature. The virucidal activity was neutralized by the<br />

addition <strong>of</strong> 2% Tween-80 and the samples were then mixed with helper cells and<br />

spread over the surface <strong>of</strong> agar plates using a s<strong>of</strong>t agar overlay. After 16h <strong>of</strong> incubation<br />

at 30ºC, the levels <strong>of</strong> Listeria contamination were determined by the presence <strong>of</strong><br />

phage plaques. It was demonstrated that the method was able <strong>to</strong> detect the presence<br />

<strong>of</strong> bacteria, however only in samples containing high numbers <strong>of</strong> L. monocy<strong>to</strong>genes.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Mapping <strong>of</strong> molecular pr<strong>of</strong>iles associated <strong>to</strong> Listeria monocy<strong>to</strong>genes<br />

food isolates circulating in Italy<br />

De Cesare, A. 1, Parisi, A. 2, La<strong>to</strong>rre, L. 2 and Manfreda, G. 1<br />

1. Alma Mater Studiorum – University <strong>of</strong> Bologna, Italy<br />

2. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale della Puglia e Basilicata, Italy<br />

Listeria monocy<strong>to</strong>genes is a bacterial pathogen transmitted <strong>to</strong> humans from a variety<br />

<strong>of</strong> foods. Although most human listeriosis infections in Italy are being generally<br />

considered <strong>to</strong> represent sporadic cases, recent results suggest that a considerable<br />

number <strong>of</strong> listeriosis might occur in clusters, many <strong>of</strong> which could<br />

represent single-source outbreaks. To support tracing <strong>of</strong> food sources <strong>of</strong> human<br />

listeriosis infections, a long term mapping <strong>of</strong> serotypes, ribotypes and AFLP types<br />

associated <strong>to</strong> Listeria monocy<strong>to</strong>gens isolates collected from different meat and<br />

cheese products as well as food production plant environments has been performed.<br />

The results achieved show that AFLP pr<strong>of</strong>iles <strong>of</strong> Listeria monocy<strong>to</strong>gens<br />

isolates from the same food source cluster <strong>to</strong>gether with a similarity ≥ 84 %,<br />

w<strong>here</strong>as AFLP clusters grouping isolates from different sources show a similarity<br />

ranging between 60 and 78 %. Isolates belonging <strong>to</strong> the same AFLP cluster show<br />

common ribotypes and serotypes. In particular, isolates from meat based products,<br />

such as fresh and fermented sausages, are mainly classified within EcoRI ribotypes<br />

14 S2 and 202 S2 and serotypes 1/2c and 1/2b. Furthermore, cheese products<br />

are classified within the EcoRI ribotype 204 S5 and serotype 1/2a.<br />

Zoonotic aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes isolates<br />

from zebu dairy animals<br />

Parihar, V. S. 1,3, Barbuddhe, S. B. 1, Kalorey, D. R. 2, Kotwal, S. 4,Danielsson Tham, M.-L. 3<br />

and Tham, W. 3<br />

1. ICAR Research Complex for Goa, Ela, Old Goa, India<br />

2. Department <strong>of</strong> Microbiology, Maharashtra Animal and Fishery Sciences University, Nagpur, India<br />

3. Örebro University, Sweden<br />

4. Sher-e- Kashmir University <strong>of</strong> Agricultural Sciences & Technology <strong>of</strong> Jammu, J&K<br />

Listeria monocy<strong>to</strong>genes is a non acid-fast, Gram-positive pathogen, which is considered<br />

food- and feed-borne. Knowledge <strong>of</strong> the direct and indirect transmission <strong>of</strong><br />

L. monocy<strong>to</strong>genes between animals and humans, is however, limited. The aim <strong>of</strong><br />

the current study was <strong>to</strong> characterize isolates <strong>of</strong> L. monocy<strong>to</strong>genes obtained from<br />

zebu dairy animals with mastitis. Fifteen isolates <strong>of</strong> L. monocy<strong>to</strong>genes from the<br />

same number <strong>of</strong> clinical and subclinical cases <strong>of</strong> mastitis were obtained from five<br />

farms harbouring al<strong>to</strong>gether 90 zebu dairy animals. The isolates were characterized<br />

by serotyping, multiplex PCR (plcA, hlyA, actA and iap genes) and pulsedfield<br />

gel electrophoresis (PFGE) using Asc Ι and Apa Ι enzymes. All the isolates<br />

characterized belonged <strong>to</strong> serovar 4b and exhibited four virulence associated genes<br />

<strong>of</strong> L. monocy<strong>to</strong>genes. PFGE revealed that all isolates belonged <strong>to</strong> the the same<br />

pulsovar. One <strong>of</strong> the reasons that all isolates belonged <strong>to</strong> the same pulsovar despite<br />

being from different farms may be that certain types <strong>of</strong> L. monocy<strong>to</strong>genes are<br />

adapted <strong>to</strong> specific niches. The presence <strong>of</strong> an identical pulsovar could also indicate<br />

a common source <strong>of</strong> infection for the different farms. All the farms delivered<br />

milk <strong>to</strong> the same dairy. The same pulsovar <strong>of</strong> L. monocy<strong>to</strong>genes was isolated from<br />

zebu dairy animals with mastitis from five different farms.<br />

R E F E R E N C E<br />

C / P<br />

116<br />

R E F E R E N C E<br />

C / P<br />

117


R E F E R E N C E<br />

C / P<br />

118<br />

R E F E R E N C E<br />

C / P<br />

119<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Performance <strong>of</strong> ALOA and Palcam agars for detection <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes in naturally contaminated raw meat products<br />

Grava<strong>to</strong> Rowlands, R. E. 1, Asturiano Ris<strong>to</strong>ri, C., Geraldes Martins, C. 1,<br />

Jakabi, M. 1 and Gombossy de Melo Franco, B. D. 2<br />

1. IAL, Brazil<br />

2. USP, Brazil<br />

The correct isolation and identification <strong>of</strong> Listeria monocy<strong>to</strong>genes in foods heavily<br />

contaminated with competing microorganisms may be difficult, especially when<br />

L. monocy<strong>to</strong>genes is present in low numbers. ALOA (Agar Listeria Ottaviani &<br />

Agosti) is a selective chromogenic agar that minimises the growth <strong>of</strong> contaminating<br />

organisms, and Listeria spp. grow on this medium producing blue/green<br />

colonies, with pathogenic species producing a characteristic opaque halo after 24<br />

h at 37 °C. Non-Listeria spp. produce white colonies. The aim <strong>of</strong> this study was <strong>to</strong><br />

evaluate the performance <strong>of</strong> ALOA for investigation <strong>of</strong> L. monocy<strong>to</strong>genes in 552<br />

samples <strong>of</strong> naturally contaminated raw meat products, comparing the results <strong>to</strong><br />

those obtained using the conventional Palcam agar. Meat samples were collected in<br />

supermarkets in the city <strong>of</strong> Sao Paulo, SP, Brazil, and comprised 138 beef sausages,<br />

138 pork sausages, 138 ground beef samples and 138 chicken leg samples. Using<br />

the ISO 11290-1:1996/Amd.1:2004 detection method, L. monocy<strong>to</strong>genes was present<br />

in 234 (42.4 %) samples. ALOA was more effective than Palcam agar for isolation <strong>of</strong><br />

L. monocy<strong>to</strong>genes (p ≤ 0.05): 82 samples (35 %) were positive in ALOA only, 23<br />

(9.8 %) in Palcam agar only, and 129 (55.1 %) samples were positive in the two<br />

media simultaneously. When the ISO 11290-2:1998 enumeration method was used<br />

95 samples (17.2 %) were positive for L. monocy<strong>to</strong>genes, and among them, 31 samples<br />

(32.6 %) were positive in ALOA only, 21 (22.1 %) in Palcam agar only, and 43<br />

(45.3 %) samples were positive in the two media simultaneously. However, counts<br />

<strong>of</strong> L. monocy<strong>to</strong>genes using the two media did not differ significantly (p > 0.05). Results<br />

<strong>of</strong> this survey indicate that the use <strong>of</strong> ALOA both for detection and enumeration<br />

<strong>of</strong> L. monocy<strong>to</strong>genes increases the effectiveness <strong>of</strong> the two ISO methods when<br />

raw meat products are tested.<br />

Comparison <strong>of</strong> MOPS-BLEB and Fraser as secondary enrichment<br />

broths for Listeria monocy<strong>to</strong>genes<br />

Upham, J., Huszczynski, G., Mosher, M., Borza, A., Dorey, M., Bosley, J., Hara, K.,<br />

Mutanda, C., Liu, J., Byrne, B. and Douey, D.<br />

Canadian Food Inspection Agency<br />

The two methods used by CFIA labora<strong>to</strong>ries for detection <strong>of</strong> L.monocy<strong>to</strong>genes from<br />

foods employ bipartite selective enrichment procedures <strong>to</strong> increase pathogen concentration<br />

prior <strong>to</strong> detection. The methods use the same bacteriological media for<br />

primary enrichment but different secondary enrichment media. Secondary media<br />

MOPS-Buffered Listeria Enrichment Broth (MOPS-BLEB) and Fraser broth were<br />

compared for their ability <strong>to</strong> facilitate growth <strong>of</strong> L. monocy<strong>to</strong>genes. Foods contaminated<br />

with Listeria were subjected <strong>to</strong> a common primary enrichment followed by<br />

parallel secondary enrichment in MOPS-BLEB and Fraser broth. Listeria was isolated<br />

and enumerated by plating the broths on selective agar. Of 164 food samples,<br />

presumptive L.monocy<strong>to</strong>genes colonies were isolated from 164 using MOPS-BLEB<br />

and 162 using Fraser broth, demonstrating broth equivalence (p = 0.50, Chisquare).<br />

Quantitative analysis revealed significantly higher levels <strong>of</strong> L.monocy<strong>to</strong>genes<br />

in MOPS-BLEB than Fraser broth (log10 CFU, 8.88 and 7.99, p < 0.0001 by<br />

paired t test). Interestingly, for foods inoculated with a mixture <strong>of</strong> L.monocy<strong>to</strong>genes<br />

and a non-pathogenic species <strong>of</strong> Listeria, L.innocua, concentrations <strong>of</strong> the<br />

later were not significantly different in MOPS-BLEB compared <strong>to</strong> Fraser broth<br />

(log10 CFU, 8.85 and 8.47, p = 0.06), resulting in a higher ratio <strong>of</strong> L.monocy<strong>to</strong>genes<br />

<strong>to</strong> L.innocua in MOPS-BLEB. Enrichment broths used by CFIA were found <strong>to</strong> be<br />

equivalent for enabling cultural detection <strong>of</strong> L. monocy<strong>to</strong>genes, however, MOPS-<br />

BLEB produced significantly higher pathogen concentrations and pathogen <strong>to</strong><br />

competi<strong>to</strong>r ratios and thus has potential <strong>to</strong> enable earlier detection and facilitate<br />

detection when competi<strong>to</strong>rs are present.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Comparison <strong>of</strong> UVM, Palcam and Oxoid Novel Enrichment (ONE) broth<br />

as primary enrichment broths for Listeria monocy<strong>to</strong>genes<br />

Upham, J., Mosher, M., Borza, A., Huszczynski, G., Dorey, M., Eloranta, K. and Douey, D.<br />

Canadian Food Inspection Agency<br />

The purpose <strong>of</strong> this study was <strong>to</strong> compare three primary enrichment broths for<br />

ability <strong>to</strong> facilitate detection <strong>of</strong> Listeria monocy<strong>to</strong>genes in foods: 1) UVM I, 2) Palcam,<br />

and 3) a single-step enrichment broth from Oxoid Company, ONE broth.<br />

Media was directly inoculated with healthy or heat-stressed L. monocy<strong>to</strong>genes and<br />

incubated for up <strong>to</strong> 48 hr, and growth curves were constructed by diluting and plating<br />

the enrichment cultures at regular time intervals <strong>to</strong> determine cell concentration<br />

(cfu/ml). Growth curves were also prepared from media that had been used <strong>to</strong><br />

enrich various food products (deli meats, dairy products, fresh produce and<br />

seafood) that were artificially contaminated with L. monocy<strong>to</strong>genes alone, or L.<br />

monocy<strong>to</strong>genes in the presence <strong>of</strong> the background organisms Escherichia coli and<br />

Staphylococcus epidermidis, or in the presence <strong>of</strong> a non-pathogenic Listeria species,<br />

L. innocua. In the absence <strong>of</strong> food matrices or competitive microorganisms, the lag<br />

phase <strong>of</strong> heat-stressed L. monocy<strong>to</strong>genes was significantly shorter in Palcam broth<br />

(6.5h) compared <strong>to</strong> UVM and ONE broth (each 7.8h), and the exponential growth<br />

rate was significantly greater in Palcam broth (0.371) and ONE broth (0.396) compared<br />

<strong>to</strong> UVM broth (0.277). Similar differences in growth rates were observed<br />

when broths were used <strong>to</strong> enrich L. monocy<strong>to</strong>genes-contaminated food products.<br />

When food were contaminated with a equal mixture <strong>of</strong> L. monocy<strong>to</strong>genes and L. innocua,<br />

the exponential growth rate <strong>of</strong> L. innocua was notably faster than L. monocy<strong>to</strong>genes<br />

in UVM and Palcam broth, w<strong>here</strong>as the opposite was true in ONE broth.<br />

In addition, the presence <strong>of</strong> L. innocua caused a decrease in L. monocy<strong>to</strong>genes<br />

growth rate in Palcam broth and <strong>to</strong> a lesser extent in UVM broth, but not in ONE<br />

broth. These studies suggest that ONE broth may be preferable <strong>to</strong> UVM and Palcam<br />

broths <strong>to</strong> facilitate detection <strong>of</strong> L. monocy<strong>to</strong>genes when L. innocua is present.<br />

Isolation and characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes from asazuke<br />

(Japanese light pickles)<br />

Maklon, K., Kusumo<strong>to</strong>, A., Makino, S.-I. and Kawamo<strong>to</strong>, K.<br />

Obihiro Univ. Agri. Vet. Med., Japan<br />

Asazuke is a ready-<strong>to</strong>-eat Japanese light pickle, mainly made <strong>of</strong> vegetables which<br />

are known <strong>to</strong> be one <strong>of</strong> the sources <strong>of</strong> Listeria monocy<strong>to</strong>genes contamination. Although<br />

asazuke is a popular side dish in Japan, the hazard <strong>of</strong> bacterial contamination<br />

has not been evaluated yet. In this study, we investigated the prevalence <strong>of</strong> L.<br />

monocy<strong>to</strong>genes, Salmonella spp., Escherichia coli O157 and coliforms in 108 asazuke<br />

samples that randomly collected from supermarkets in Obihiro (Hokkaido prefecture,<br />

Japan) during the period <strong>of</strong> June <strong>to</strong> November 2007. Twelve (11.1 %) L.<br />

monocy<strong>to</strong>genes were isolated with predominant serotype 4b (7 isolates) followed by<br />

1/2a (2 isolates), 1/2b, 3b and 4c (1 isolate each) while Salmonella spp., E. coli O157<br />

and coliforms were not detected. All L. monocy<strong>to</strong>genes isolates demonstrated hemolytic<br />

activity by CAMP test and possessed all the virulence-associated genes<br />

(prfA, actA, mpl, inlA, inlC, plcA, plcB, hly, iap, clpC and opuCA) by PCR, those revealed<br />

their potential pathogenicity. Moreover, 7 out <strong>of</strong> 12 isolates were from<br />

asazuke produced by one fac<strong>to</strong>ry and pulsed-field gel electrophoresis (PFGE) pr<strong>of</strong>iles<br />

suggested that 6 <strong>of</strong> them were indistinguishable. Additional investigation <strong>of</strong> L.<br />

monocy<strong>to</strong>genes contamination was performed in the asazuke fac<strong>to</strong>ry environment<br />

and 23 out <strong>of</strong> 60 swabs (38.33 %) were isolated. Comparison <strong>of</strong> PFGE pr<strong>of</strong>iles<br />

showed relatedness between food and environmental isolates. Taken <strong>to</strong>gether, our<br />

results indicated that asazuke was contaminated with L. monocy<strong>to</strong>genes, which<br />

was probably from vegetables, human sources or from production lines. Further<br />

analysis <strong>to</strong> clarify the source <strong>of</strong> pathogen contamination is needed <strong>to</strong> establish a<br />

prevention method and consumers should be aware <strong>of</strong> asazuke consumption that<br />

might be contaminated with L. monocy<strong>to</strong>genes.<br />

R E F E R E N C E<br />

C / P<br />

120<br />

R E F E R E N C E<br />

C / P<br />

121


R E F E R E N C E<br />

C / P<br />

122<br />

R E F E R E N C E<br />

C / P<br />

123<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Grouping <strong>of</strong> human Listeria monocy<strong>to</strong>genes isolates<br />

Lopez-Valladares, G., Danielsson-Tham, M.-L. and Tham, W.<br />

School <strong>of</strong> Hospitality, Culinary Arts & Meal Sciences, Örebro University, Grythyttan, Sweden<br />

More than 800 clinical isolates <strong>of</strong> Listeria monocy<strong>to</strong>genes from humans in Sweden<br />

were characterized by serotyping and pulsed-field gel electrophoresis. The restriction<br />

enzyme used was AscI. The serovar 4b isolates were divided in<strong>to</strong> 27 PFGE<br />

types, the serovar 1/2a and 1/2c isolates in<strong>to</strong> 108, and the serovar 1/2b and 3b isolates<br />

in<strong>to</strong> 24. PFGE types with identical pr<strong>of</strong>iles <strong>of</strong> small bands (< 145 kbp) were<br />

grouped <strong>to</strong>gether with PFGE designations <strong>of</strong> A-Z. This grouping was based on the<br />

assumption DNA rearrangements involving insertions, deletions, or transpositions<br />

<strong>of</strong> genes are less common in the small fragments than in large fragments.<br />

The 27 serovar 4b PFGE types were further divided in<strong>to</strong> 5 groups and the 108<br />

serovar 1/2a and 1/2c PFGE types in<strong>to</strong> 16 groups. The serovar 1/2b and 3b types<br />

(n = 24) were divided in<strong>to</strong> five groups. Part <strong>of</strong> the collection <strong>of</strong> L. monocy<strong>to</strong>genes<br />

isolates was analyzed with multiple-locus variable-number tandem-repeat<br />

analysis and the clusters were generally in good agreement with both the clusters<br />

generated by PFGE and serotyping data. The advantage <strong>of</strong> dividing L. monocy<strong>to</strong>genes<br />

PFGE types in<strong>to</strong> groups is that the pr<strong>of</strong>ile <strong>of</strong> every new isolate characterised<br />

with PFGE can easily and quickly be identified by studying the group affiliation <strong>of</strong><br />

the isolate and then identifying the matching type within the group.<br />

Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated from food<br />

and environmental samples<br />

Nucera, D. M. 1, Lomonaco, S. 1, Manila Bianchi, D. 2, Decastelli, L. 2, Grassi, M. A. 1 and Civera, T. 1<br />

1. Università degli Studi di Torino, Italy<br />

2. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico del Piemonte Liguria e Valle d’Aosta, Italy<br />

This study aimed <strong>to</strong> investigate type diversity and distribution over time and<br />

across sources by subtyping via Pulsed Field Gel Electrophoresis (PFGE) a set <strong>of</strong><br />

300 L. monocy<strong>to</strong>genes strains collected over a five year period. Five clinical human<br />

strains isolated in the same geographic area and period <strong>of</strong> the study were also included<br />

in the analysis. The isolated strains belonged <strong>to</strong> serotype 1/2a (45 % <strong>of</strong> the<br />

samples), 1/2c (22 %), 4b/4e (16 %); however, 5 % <strong>of</strong> the strains were untypeable.<br />

Significant associations were observed between serotype 1/2a with dairy (O.R.=<br />

13.9; χ 2 p < 0.05) and 1/2c with meat (O.R.= 33.3; χ 2 p < 0.05). PFGE results highlighted<br />

152 pulsotypes shared among two or more samples: 9 pulsotypes were recurrent<br />

and 6 were shared between strains isolated from different food and environmental<br />

sources. The other generated pulsotypes were source specific and/or<br />

retrieved in one year only. These findings may indicate the presence <strong>of</strong> both nicheadapted<br />

as well as ubiqui<strong>to</strong>us PFGE types. Moreover, no PFGE types were shared<br />

between strains collected from food/environmental isolates and human clinical<br />

strains. The results <strong>of</strong> this research underline and confirm the importance <strong>of</strong> implementing<br />

a broad typing database which could facilitate epidemiological investigations<br />

and enhance novel food attribution modelling approaches for the identification<br />

<strong>of</strong> listeriosis outbreaks and the related sources.These data, even if focused<br />

on strains collected in a limited geographic area, may pose the ground <strong>to</strong> evidence<br />

how large subtype databases may aid in the detection <strong>of</strong> common- and source specific-<br />

types. However, more comprehensive databases, both at national and at European<br />

level, are needed <strong>to</strong> reveal L. monocy<strong>to</strong>genes strains diversity and <strong>to</strong> provide<br />

useful data for epidemiological investigations.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Development <strong>of</strong> a multiplex SNP-typing method <strong>to</strong> identify the four<br />

epidemic clones <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

Lomonaco, S. 1, Civera, T. 1, Dalmasso, A. 1, Knabel, S. J. 2 and Bottero, M. T. 1<br />

1. Università degli Studi di Torino, Italy<br />

2. The Pennsylvania State University, USA<br />

The concept <strong>of</strong> epidemic clone (EC) is used <strong>to</strong> describe closely related strains <strong>of</strong> L.<br />

monocy<strong>to</strong>genes belonging <strong>to</strong> geographically and temporally distinct outbreaks. Currently,<br />

four ECs have been identified: ECI, ECII, ECIII, and ECIV (also ECIa). The<br />

current classification <strong>of</strong> ECs is based on results from different subtyping techniques<br />

such as RFLP analysis, hybridization assays, PFGE and ribotyping. Recently, a multiplex<br />

PCR able <strong>to</strong> identify ECI, ECII and ECIII was developed. In addition, sequence-based<br />

methods, such as multivirulence-locus sequence typing, were able<br />

<strong>to</strong> further subtype EC strains and also <strong>to</strong> highlight potentially informative single<br />

nucleotide polymorphisms (SNPs). The aim <strong>of</strong> the present study was <strong>to</strong> develop an<br />

alternative method for the identification <strong>of</strong> all four ECs by developing a SNP-typing<br />

method based on minisequencing. This technique simultaneously analyses several<br />

SNPs based on a Primer-Extension Reaction (PER). Extension primers are designed<br />

immediately adjacent <strong>to</strong> the SNPs and extended by a single [F]ddNTP. Results are<br />

then visualized with a genetic analyzer as specific patterns. Six SNPs, yielding a<br />

specific pattern for each EC, were selected. These SNPs were located on 4 virulence<br />

genes (inlA, inlJ, inlB and srtA). Gene fragments were preliminary amplified with a<br />

multiplex PCR and then used as template for the PER reaction. Analyzed samples<br />

were 39 ECs strains, 25 strains from the ILSI Diversity Subset and 22 non ECstrains.<br />

All ECs gave a clone-specific pattern. Moreover, 9 strains not previously<br />

classified as ECs showed the same pr<strong>of</strong>ile as ECI, ECII and ECIV. These findings<br />

were confirmed with other biomolecular tests (e.g. ECs-specific PCR) and by considering<br />

available epidemiological data for these strains. The proposed SNP-typing<br />

assay is potentially high-throughput and amenable <strong>to</strong> au<strong>to</strong>mation, allowing a<br />

rapid identification <strong>of</strong> all four ECs and can t<strong>here</strong>fore be used as a screening method<br />

in the analysis <strong>of</strong> L. monocy<strong>to</strong>genes strains.<br />

Listeria monocy<strong>to</strong>genes incidents reported <strong>to</strong> the UK Food Standards<br />

Agency from 2000 <strong>to</strong> 2008<br />

Aish, J.<br />

Food Standards Agency, UK<br />

The UK Food Standards Agency contributes <strong>to</strong> the management <strong>of</strong> a wide range <strong>of</strong><br />

food related incidents involving microbiological, chemical, radiological and physical<br />

hazards. One important foodborne microbiological hazard for which the<br />

Agency oversees incident investigations is Listeria monocy<strong>to</strong>genes. When incidents<br />

are reported <strong>to</strong> the Agency the relevant data is recorded on the Agency’s Incidents<br />

Database. Notification is received from a wide range <strong>of</strong> food businesses, Local Authorities,<br />

the Health Protection Agency and other Government Departments. Information<br />

held on the database was analysed <strong>to</strong> determine the number <strong>of</strong> incidents<br />

involving Listeria spp. that were reported <strong>to</strong> the Agency between 2000 and<br />

2008 and <strong>to</strong> identify that main food categories that were involved. During the nine<br />

year period from 2000 <strong>to</strong> 2008 approximately 140 incidents caused by the contamination<br />

<strong>of</strong> food with Listeria spp. were reported <strong>to</strong> the Agency, the majority <strong>of</strong><br />

which involved L. monocy<strong>to</strong>genes. Most <strong>of</strong> the incidents involved ready-<strong>to</strong>-eat<br />

meat/meat products, cheese, fish/shellfish and sandwiches/sandwich fillings. L.<br />

monocy<strong>to</strong>genes is an important foodborne pathogen. Although healthy people are<br />

not usually at risk <strong>of</strong> illness, listeriosis can be an issue for vulnerable groups such<br />

as pregnant women and the immunocompromised. In addition <strong>to</strong> overseeing incident<br />

investigations the Agency’s 4C’s strategy (cleaning, cooking, chilling and<br />

avoiding cross-contamination) is important in reducing the risk from L. monocy<strong>to</strong>genes.<br />

The Agency also provides food safety advice <strong>to</strong> vulnerable groups and<br />

commissions research <strong>to</strong> further our understanding <strong>of</strong> this organism.<br />

R E F E R E N C E<br />

C / P<br />

124<br />

R E F E R E N C E<br />

C / P<br />

125


R E F E R E N C E<br />

C / P<br />

183<br />

R E F E R E N C E<br />

C / P<br />

184<br />

AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Genetic diversity <strong>of</strong> Listeria monocy<strong>to</strong>genes in broiler flocks<br />

Courtillon, C. Toquin, M.-T., Le Nôtre, Y., Fravalo, P. and Mansour Chemaly, M.<br />

AFSSA, France<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogen, persistent in industrial environment<br />

and is responsible <strong>of</strong> listeriosis, an infectious and rare disease. The aim <strong>of</strong><br />

this study was <strong>to</strong> update and create data set from broiler flocks regarding their<br />

contamination by Listeria monocy<strong>to</strong>genes. 145 flocks have been sampled, 85 conventional<br />

and 60 free range flocks. Molecular characterization by PFGE with two<br />

restriction enzymes ApaI and AscI was performed in order <strong>to</strong> establish the clonal<br />

relationships <strong>of</strong> the isolates and <strong>to</strong> trace the contamination between the flocks. A<br />

<strong>to</strong>tal <strong>of</strong> 126 isolates were genotyped. The prevalence in broilers has been estimated<br />

at 31.7 % (46 positive samples on 145). The serotyping revealed the presence <strong>of</strong><br />

serotype 1/2a in majority: 50 % vs. 31.7 % for serogroup 4 (4e and 4b) and 18.3 % for<br />

1/2b. Serotype 1/2a is the most present in conventional farms (37.3 %) contrary <strong>to</strong><br />

free range farms (12.7%) w<strong>here</strong> serogroup 4 is dominant (23 % vs. 8.7 %). PFGE<br />

typing showed a high discrimina<strong>to</strong>ry index: 0.971 for both combined enzymes. 47<br />

genetic patterns were identified. Dendrograms, obtained with BioNumerics s<strong>of</strong>tware,<br />

confirmed the high genetic variability between isolates (55.7 % <strong>of</strong> similarity)<br />

especially in serotype 1/2a (65.5 %). This index (0.954) was similar <strong>to</strong> the one<br />

obtained when studying laying hen flocks. The serogroup 4 and the serotype 1/2b<br />

seemed more clonal, 0.892 and 0.798, respectively. Regarding the production system,<br />

results between conventional and free range farms are comparable, 0.959 and<br />

0.948, respectively. The PFGE showed that two identical pr<strong>of</strong>iles can be found in<br />

different flocks and different geographical locations while in a same flock, patterns<br />

can be very different. The breeding type is not linked <strong>to</strong> bacteria genetic<br />

identity as are serotypes. In conclusion, this investigation brought new and updated<br />

data regarding Listeria monocy<strong>to</strong>genes in broiler flocks which were not available<br />

before.<br />

Tracing Listeria monocy<strong>to</strong>genes contaminations throughout<br />

the processing chain <strong>of</strong> a typical Italian pork meat product<br />

using Pulsed Field Gel Electrophoresis (PFGE) characterization<br />

Annunziata Prencipe, V., Acciari, V., Torresi, M., Migliorati, G.,<br />

Marfoglia, C. and Valentina Rizzi, V.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

Pulsed Field Gel Electrophoresis (PFGE) and serotyping analyses were performed<br />

on 124 Listeria monocy<strong>to</strong>genes strains isolated from a prevalence study throughout<br />

the Parma dry-cured ham production, in order <strong>to</strong> evaluate contamination patterns.<br />

Strains had been isolated at three different points <strong>of</strong> the production process:<br />

I) fecal samples/carcasses in slaughterhouses; II) fresh hams after cutting; III)<br />

de-boned, dry-cured and packaged hams at the end <strong>of</strong> the production chain. Traceability<br />

<strong>to</strong> corresponding feces, fresh hams and dry-cured hams had been assured<br />

for all carcasses, so every strain was tested <strong>to</strong> evaluate the presence <strong>of</strong> contaminations<br />

in previous or next phases <strong>of</strong> the production process. Genetic relationships <strong>of</strong><br />

56 different pulsotypes obtained from the combination <strong>of</strong> AscI and ApaI macrorestriction<br />

patterns were graphically represented and analyzed. Six different clusters<br />

were identified. The analysis <strong>of</strong> PFGE patterns obtained from the same sample<br />

examined at the three different points <strong>of</strong> the production process (carcass – fresh<br />

ham – cured ham) did not produce identical or highly similar pulsotypes. The only<br />

pulsotype occurring in a strain from faeces was not recovered from other materials.<br />

Strains from the same production plants were usually genetically homogeneous<br />

or highly similar. In the plants w<strong>here</strong> t<strong>here</strong> was not genetic homogeneity,<br />

one pulsotype or cluster was always sharply prevalent over the others. The presence<br />

<strong>of</strong> single or prevalent pulsotypes in the same processing plants could be associated<br />

<strong>to</strong> persistent strains <strong>of</strong> Listeria monocy<strong>to</strong>genes. This is corroborated by<br />

the highest prevalence <strong>of</strong> 1/2c serotype strains, which ad<strong>here</strong> significantly more <strong>to</strong><br />

working surfaces, isolated from all stages <strong>of</strong> the process. Genetically homogeneous<br />

or highly similar strains were also recovered in different establishments, at different<br />

phases <strong>of</strong> the production process. This situation could be bounded <strong>to</strong> previous<br />

sporadic contaminations between different production stages, which could<br />

have allowed the localization <strong>of</strong> strains in<strong>to</strong> favorable niches inside the processing<br />

environment.


AREA // C // Epidemiologic and clinical aspects <strong>of</strong> Listeria monocy<strong>to</strong>genes and listeriosis<br />

POSTER PRESENTATIONS // P / 82–125, 183–186<br />

Production and validation <strong>of</strong> capture-ELISA kit based on monoclonal<br />

antibodies specific for Listeria monocy<strong>to</strong>genes in foodstuffs<br />

Portanti, O., Di Febo, T., Luciani, M., Pompilii, C., Armillotta, G., Principe, V.,<br />

Lelli, R. and Semprini, P.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

A Listeria monocy<strong>to</strong>genes ELISA test was produced for use in the screening <strong>of</strong><br />

foods. Validation was carried out according <strong>to</strong> the criteria for qualitative analysis<br />

set out in ISO 16140:2003. The ELISA kit has satisfied the norm requirements in<br />

terms <strong>of</strong> relative accuracy, relative sensitivity, relative specificity, relative detection<br />

level, inclusivity and exclusivity. The kit is a capture-ELISA test applied <strong>to</strong><br />

the sample enrichment broth, based both on a monoclonal antibody (9B8F7) <strong>to</strong><br />

L. monocy<strong>to</strong>genes and a peroxidase-conjugated anti-Listeria monoclonal antibody<br />

(6F12C8). The method has been validated in meat, fish and dairy products by<br />

means <strong>of</strong> incurred and artificially contaminated samples with target and non target<br />

strains. Results for relative accuracy, relative sensitivity, relative specificity<br />

and relative detection levels are in agreement with ISO 11290-1:1996 requirements.<br />

Capture-ELISA test is able <strong>to</strong> detect Listeria with 100 % inclusivity and exclusivity<br />

in the above mentioned matrices. Capture-ELISA kit was produced in a ready-<strong>to</strong>use<br />

format and can be adopted by microbiological labs that need fast analytical<br />

methods. It also shows the same features as the internationally recognised ISO<br />

11290 techniques, as required by UNI EN ISO 17025:2005.<br />

Production <strong>of</strong> a reference material for microbiological tests<br />

containing Listeria innocua<br />

Pomilio, F., Ricci, L, Di Giannatale, E., Semprini, P., Candeloro, L. and Migliorati, G.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

The study describes the production <strong>of</strong> a reference material constituted by glass<br />

balls, covered by skimmed milk powder micronized and artificially contaminated<br />

with Listeria innocua (ATCC 33090 strain). The lot <strong>of</strong> prepared material, constituted<br />

by 1g unities <strong>of</strong> balls, has been made in test-tubes, submitted <strong>to</strong> a stability<br />

and homogeneity evaluation through the determination <strong>of</strong> the mesophilic bacterial<br />

count and following Listeria innocua identification. The material stability, kept<br />

at –20 °C, has been evaluated at 54, 180 and 355 days. Data pointed out the decreasing<br />

tendency <strong>of</strong> the bacterial count, statistically significant, at 54 days. At 180<br />

and 355 days, the lot resulted stable. Regression analysis at 54 days (t-value -2,67)<br />

has underlined a significant decreasing trend (p-value < 0.05). In the observed period<br />

the decrease has been equal <strong>to</strong> 8.75 UFC/g. Samples examined in the same<br />

time lapses resulted homogeneous in the interval between 180 and 355 days. Only<br />

a sample out <strong>of</strong> the 105 examined resulted negative. After 355 days <strong>of</strong> preservation<br />

at –20 °C, the challenge test has been carried out at 5 different temperatures<br />

(-20 °C; +4 °C; +22 °C; +30 °C; +37 °C). The material evaluated with non-parametric<br />

tests, for independent samples resulted stable at -20 °C and +4 °C. Furthermore, for<br />

the reference value definition, a pr<strong>of</strong>iciency test has been organized, which allowed<br />

<strong>to</strong> assign the value <strong>of</strong> 3.2 UFC/g <strong>to</strong> the material. In the same study, 35 samples<br />

out <strong>of</strong> 244 (15.4 %) resulted negative. Used methodology could be applied <strong>to</strong><br />

reference materials production also on other micro-organisms. The adopted technology<br />

has the benefit <strong>of</strong> operating in “closed” systems, making bio-safety <strong>of</strong> opera<strong>to</strong>rs<br />

and environment easily manageable.<br />

R E F E R E N C E<br />

C / P<br />

185<br />

R E F E R E N C E<br />

C / P<br />

186


R E F E R E N C E<br />

D / P<br />

126<br />

R E F E R E N C E<br />

D / P<br />

127<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Colonization <strong>of</strong> a newly constructed commercial chicken further<br />

processing plant with Listeria monocy<strong>to</strong>genes<br />

Berrang, M. E. 1, Meinersmann, R. 1, Frank, J. 2 and Ladely, S. 1<br />

1. USDA-Agricultural Research Service, USA<br />

2. University <strong>of</strong> Geórgia, USA<br />

This study was undertaken <strong>to</strong> determine potential sources <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in a newly constructed chicken further processing plant and document the<br />

eventual colonization <strong>of</strong> the facility by this pathogen. To ascertain the colonization<br />

status <strong>of</strong> the plant, floor drains were sampled after a production shift and again<br />

after a clean up shift on a roughly monthly basis for 21 months. Potential sources <strong>of</strong><br />

L. monocy<strong>to</strong>genes <strong>to</strong> the plant included: incoming raw meat, incoming fresh air<br />

and personnel. Nearby environment and community samples were also examined.<br />

All L. monocy<strong>to</strong>genes detected were subjected <strong>to</strong> DNA sequence based subtyping.<br />

L. monocy<strong>to</strong>genes was not detected in the plant before the commencement <strong>of</strong> processing<br />

operations. Within four months, several subtypes <strong>of</strong> L. monocy<strong>to</strong>genes were<br />

detected in floor drains both before and after cleaning and sanitizing operations.<br />

No L. monocy<strong>to</strong>genes was detected on filters for incoming air, samples associated<br />

with plant employees, or a nearby discount shopping center. One subtype <strong>of</strong> L.<br />

monocy<strong>to</strong>genes was detected in a natural stream near the plant however, this subtype<br />

was never detected inside the plant. Eight subtypes <strong>of</strong> L. monocy<strong>to</strong>genes were<br />

detected in raw meat staged for further processing; one <strong>of</strong> the raw meat subtypes<br />

was indistinguishable from a persistent drain subtype recovered after cleaning on<br />

eight occasions in four different drains. Poultry further processing plants are likely<br />

<strong>to</strong> become colonized with L. monocy<strong>to</strong>genes; raw product is an important source <strong>of</strong><br />

the organism <strong>to</strong> the plant.<br />

Episcopic differential interference contrast/epifluorescence<br />

microscopy <strong>to</strong> characterise in situ Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms<br />

on stainless steel surfaces<br />

Gião, M. S. and Keevil, C. W.<br />

University <strong>of</strong> Southamp<strong>to</strong>n, UK<br />

Listeria monocy<strong>to</strong>genes is a Gram-positive pathogen associated with the contamination<br />

<strong>of</strong> several food products, including meat, dairy products and vegetables. Although<br />

this pathogen can be found in natural environments the contamination <strong>of</strong><br />

processed food can occur after contact with contaminated surfaces. Previous studies<br />

routinely focused on bi<strong>of</strong>ilm recovery and culture quantification. In situ studies<br />

utilised mostly SEM which is tedious, expensive and known <strong>to</strong> introduce artefacts.<br />

By contrast episcopic differential interference contraste/epifluorescence<br />

(EDIC/EF) microscopy is a novel and rapid light microscopy technique ideally<br />

suited <strong>to</strong> study bi<strong>of</strong>ilms in a native state. The aim <strong>of</strong> this work was <strong>to</strong> utilise the advances<br />

in EDIC/EF microscopy <strong>to</strong> study the formation <strong>of</strong> L. monocy<strong>to</strong>genes<br />

bi<strong>of</strong>ilms under different nutrient and temperature conditions. L. monocy<strong>to</strong>genes<br />

NCTC 13372 was suspended in Brain Heart Infusion (BHI), Tryp<strong>to</strong>ne Soya Broth<br />

(TSB), TSB supplemented with 0.6 % <strong>of</strong> Yeast Extract (TSBYE) and filter-sterilised<br />

tap water. Five ml <strong>of</strong> each suspension were transferred for 6 well plates containing<br />

1 cm 2 stainless steel coupons and incubated at 4 °C, 22 °C, 30 °C and 37 °C. After 4<br />

and 24 hours 2 coupons were removed for each tested medium. One coupon was<br />

stained in situ with SYTO 9 and observed under EDIC/EF microscopy and the<br />

other coupon was scraped and <strong>to</strong>tal and cultivable cells quantified by staining with<br />

SYTO 9 and plating on<strong>to</strong> BHI agar, respectively. Results showed that even for the<br />

lowest temperature, after 4 hours <strong>of</strong> incubation L. monocy<strong>to</strong>genes bi<strong>of</strong>ilms were<br />

already developing on the surfaces. Except in the case <strong>of</strong> bi<strong>of</strong>ilms formed in tap<br />

water w<strong>here</strong> cells were in general less cultivable, the different media do not influence<br />

significantly the formation <strong>of</strong> bi<strong>of</strong>ilm on stainless steel surfaces. Conversely,<br />

the in situ observation <strong>of</strong> the coupons demonstrated that the morphology <strong>of</strong> the<br />

bi<strong>of</strong>ilm is temperature dependent with implications for food processing.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Temperature dependent defect in bi<strong>of</strong>ilm formation<br />

by Listeria monocy<strong>to</strong>genes<br />

Abdalla, S. 1, Glenn, S. 1, Shama, G. 2 and Andrew, P. 1<br />

1. Department <strong>of</strong> Infection, Immunity and Inflammation, University <strong>of</strong> Leicester LE1 9HN, United<br />

Kingdom<br />

2. Department <strong>of</strong> Chemical Engineering Loughborough University, Loughborough, LE11 3TU, United<br />

Kingdom<br />

Listeria monocy<strong>to</strong>genes can cause major problems in food processing industries because<br />

<strong>of</strong> its ability <strong>to</strong> survive and grow over a wide range <strong>of</strong> environmental conditions.<br />

Furthermore L. monocy<strong>to</strong>genes can attach <strong>to</strong> and produce bi<strong>of</strong>ilm on wide<br />

variety <strong>of</strong> different surfaces in the food-processing environment, allowing the bacterium<br />

<strong>to</strong> resist <strong>to</strong> antimicrobial and sanitizing agents. The aims <strong>of</strong> this study were<br />

<strong>to</strong> investigate the ability <strong>of</strong> Listeria monocy<strong>to</strong>genes wide type strain 10403s and<br />

Tn917-LTV3 transposon mutants <strong>to</strong> attach <strong>to</strong> abiotic surface at a variety <strong>of</strong> temperatures.<br />

A published attachment assay was modified <strong>to</strong> compare the ability <strong>of</strong> the<br />

WT strain and a number <strong>of</strong> transposon mutants <strong>to</strong> attach <strong>to</strong> polystyrene surfaces<br />

over 2 hours incubation. Attached cells were detected using crystal violet staining.<br />

Certain transposon mutants were shown <strong>to</strong> have a significant reduction in the level<br />

<strong>of</strong> attachment, but this defect was temperature dependent, being evident at 30 °C<br />

and 18 °C but not at 37 °C. The identity <strong>of</strong> the transposon insertion site in these<br />

mutants is being actively pursued. It was concluded that t<strong>here</strong> is a temperature dependent<br />

involvement <strong>of</strong> some gene production in surface attachment.<br />

Mode <strong>of</strong> action <strong>of</strong> Lac<strong>to</strong>coccus lactis Sa31 antimicrobial peptide<br />

on Listeria monocy<strong>to</strong>genes ½ c<br />

Barile, M. 1, Mormile, A. 1, Ceres, C. 1, Pepe, O. 2, Cortesi, M. L. 1 and Murru, N. 1<br />

1. DISCIZIA, Faculty <strong>of</strong> Veterinary Medicine, Napoli, Italy<br />

2. Faculty <strong>of</strong> Agraria Università degli Studi di Napoli, Italy<br />

Lactic acid bacteria produce antimicrobial peptides referred <strong>to</strong> as bacteriocins<br />

that have potential as natural food preservatives. Lac<strong>to</strong>coccus lactis Sa31, isolated<br />

from gilthead breams fillets packaged in modified atmosp<strong>here</strong>s, normally produces<br />

bacteriocin in GM 17 broth with maximum activity (2.560 AU/ml) at 11 h in<br />

the late logarithmic <strong>to</strong> early stationary phase <strong>of</strong> growth. In the present work was<br />

investigated the Sa 31 ability <strong>of</strong> bacteriocin production in different liquid media,<br />

the partial characterization and mode <strong>of</strong> action <strong>of</strong> peptide against Listeria monocy<strong>to</strong>genes<br />

½ c strain. Overnight extracts <strong>of</strong> producer strain cultures in Trip<strong>to</strong>ne<br />

Soya Broth, M 17 + 0.5 % glucose (GM 17 ) and 0,5 % lac<strong>to</strong>se (LM 17 ), MRS, and BHI<br />

broth was tested against Listeria monocy<strong>to</strong>genes ½ c. To evaluate the mode <strong>of</strong> action,<br />

the crude (5.120 AU/ml -1) and partially purificated bacteriocin (10.240<br />

AU/ml -1) by precipitation with ammonium sulfate were inoculated in suspension<br />

<strong>of</strong> 50 mM potassium phospha<strong>to</strong> buffer pH 7.0 containing Listeria monocy<strong>to</strong>genes<br />

½ c (1.4 x10 8 CFU/ml). Finally, for absorption study <strong>of</strong> partial purificated bacteriocin,<br />

this was added at 10.240 AU/ml on cell <strong>of</strong> Listeria monocy<strong>to</strong>genes ½ c resuspended<br />

in 100 mM NaCl pH 7.0 incubated for 30 min. at 20 °C. The residual<br />

bacteriocin activity was tested as described above. The strain Sa 31 was able <strong>to</strong><br />

produce bacteriocin in BHI broth with maximum titre 5.120 AU/ml -1 after short<br />

incubation time. The bacteriocin, sensitive <strong>to</strong> heat and more powerful at acid pH,<br />

exhibited a bactericidal mode <strong>of</strong> action and acted rapidly. The death <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes ½ c cells was observed within 10 min. after treatment with 10.240<br />

AU/ml <strong>of</strong> bacteriocin (reduction in viable counts from 1.4 x 10 8 <strong>to</strong> 6.5x10 5 CFU/ml).<br />

The bacteriocin was completely absorbed on the sensitive strain L. monocy<strong>to</strong>genes<br />

½ c at pH 7.0. These results demonstrate the effectiveness <strong>of</strong> this bacteriocin and<br />

its producer as an inhibi<strong>to</strong>r <strong>of</strong> Listeria monocy<strong>to</strong>genes in vitro, and the possibility <strong>of</strong><br />

their applying in a food system w<strong>here</strong> post-manufacture contamination by this<br />

organism could be problematic as in fishery products as well as in other foods.<br />

R E F E R E N C E<br />

D / P<br />

128<br />

R E F E R E N C E<br />

D / P<br />

129


R E F E R E N C E<br />

D / P<br />

130<br />

R E F E R E N C E<br />

D / P<br />

131<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Tracing the source, epidemiology and persistence <strong>of</strong><br />

Listeria monocy<strong>to</strong>genes in Irish Farmhouse cheese processing facilities<br />

Jordan, K., Fox, E., O’Brien, M. and Hunt, K.<br />

Teagasc, Ireland<br />

Many outbreaks <strong>of</strong> listeriosis have been related <strong>to</strong> dairy products, although the<br />

source <strong>of</strong> contamination has rarely been determined. Sixteen farmhouse cheesemaking<br />

facilities were sampled at monthly intervals over a two-year period. Samples<br />

included raw materials, food contact surfaces, non-food contact surfaces, farm<br />

samples and product. In order <strong>to</strong> determine the source, persistence and epidemiology<br />

<strong>of</strong> the strains obtained from 13 cheesemaking facilities, isolates <strong>of</strong> L. monocy<strong>to</strong>genes<br />

were differentiated using PFGE and serotyping. At 9 <strong>of</strong> the 13 contaminated<br />

processing facilities farm samples were the source <strong>of</strong> the environmental<br />

contamination. We could not determine the source <strong>of</strong> the contamination at the<br />

remaining 4 facilities. Persistent isolates (similar isolates obtained at the same facility<br />

for > one year) were found at two cheesemaking facilities. Reduction <strong>of</strong> the<br />

occurrence <strong>of</strong> L. monocy<strong>to</strong>genes in milk could reduce the contamination <strong>of</strong> cheese<br />

processing facilities and t<strong>here</strong>fore reduce the risk <strong>of</strong> cheese contamination. Cheese<br />

producers can also take steps <strong>to</strong> reduce risk by reducing milk spillage at milk intake<br />

and sterilising any area w<strong>here</strong> milk was spilt. In addition, barriers between the<br />

farm and processing facility will help reduce the risk <strong>of</strong> contamination.<br />

Phenotypic and genotypic characteristics <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

strains isolated from a convenience food-processing plant<br />

Blatter, S., Stephan, R., Tasara, T. and Zweifel, C.<br />

Institute for Food Safety and Hygiene, University <strong>of</strong> Zurich, Switzerland<br />

L. monocy<strong>to</strong>genes as a food-borne pathogen has significant public health and economic<br />

impacts. Amongst other foods, ready-<strong>to</strong>-eat products have been implicated as<br />

routes for human infection. The aim <strong>of</strong> this study was <strong>to</strong> investigate the occurrence<br />

and genetic diversity <strong>of</strong> L. monocy<strong>to</strong>genes isolated from the environment and food<br />

products in a Swiss sandwich-producing plant. Over a 12-month period, environmental<br />

swabs (n = 2.028) collected during the working activity and after cleaning<br />

and disinfection, as well as ingredient and sandwich samples (n = 217) were collected.<br />

Swabs originated from the processing room and the equipment <strong>of</strong> five sandwich<br />

production lines. Isolation <strong>of</strong> L. monocy<strong>to</strong>genes was accomplished by culture<br />

after enrichment (Half-Fraser/Fraser broth, Palcam and Ottaviani-Agosti agar).<br />

Isolated strains were characterized by serotyping, determination <strong>of</strong> genetic lineages,<br />

Rep PCR typing, and PFGE analysis after macrorestriction. L. monocy<strong>to</strong>genes<br />

were detected in 70 (3.5 %) environmental swabs and 16 (7.4 %) samples from ingredients<br />

and sandwiches. Of the positive environmental samples, 40 % originated<br />

from three slicers, which tested repeatedly positive on several occasions. Of the 86<br />

isolated L. monocy<strong>to</strong>genes strains, 93 % belonged <strong>to</strong> serotype 1/2a and genetic lineage<br />

II, w<strong>here</strong>as the remaining strains were <strong>of</strong> serotype 1/2b and genetic lineage I.<br />

Genotyping by Rep PCR and PFGE analysis yielded each six different pr<strong>of</strong>iles. Fiftyeight<br />

(82.9 %) strains from the processing environment and 9 (56.3 %) from ingredients<br />

and sandwiches belonged <strong>to</strong> only one genotype. Strains <strong>of</strong> this genotype were<br />

found repeatedly on/in slicers, tables, conveyor belts, tables, a bread-feeding machine,<br />

spattles, air guns, salmon and egg sandwiches. Based on the genotyping results,<br />

indications <strong>of</strong> persistent contamination <strong>of</strong> the sandwich processing area with<br />

a predominant strain were evident over the whole sampling period. Moreover, in<br />

food-processing plants with low Listeria prevalence in the final product, environmental<br />

moni<strong>to</strong>ring is more effective than end product testing.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Influence <strong>of</strong> flow direction on the adhesion <strong>of</strong> Listeria monocy<strong>to</strong>genes <strong>to</strong><br />

brushed stainless steel surfaces<br />

Skovager, A. 1, Whitehead, K. 2, Ingmer, H. 3, Verran, J. 2 and Arneborg, N. 1<br />

1. Department <strong>of</strong> Food Science, Food Microbiology, Faculty <strong>of</strong> Life Sciences, University <strong>of</strong> Copenhagen,<br />

Denmark<br />

2. Division <strong>of</strong> Biology, School <strong>of</strong> Biology, Chemistry and Health Science, Faculty <strong>of</strong> Science and<br />

Engineering, John Dal<strong>to</strong>n Building, Manchester Metropolitan University, Manchester, England<br />

3. Department <strong>of</strong> Veterinary Disease Biology, Section for Microbiology, Faculty <strong>of</strong> Life sciences,<br />

University <strong>of</strong> Copenhagen, Denmark<br />

The ability <strong>of</strong> Listeria monocy<strong>to</strong>genes <strong>to</strong> form bi<strong>of</strong>ilms on the surfaces <strong>of</strong> process<br />

equipment constitutes a major health problem for the food industry. When the<br />

bacteria have succeeded in forming a bi<strong>of</strong>ilm, they are difficult <strong>to</strong> remove from<br />

the surface, due <strong>to</strong> their strong adhesion on the processing equipment. The innermost<br />

cells <strong>of</strong> a bi<strong>of</strong>ilm are also protected against the action <strong>of</strong> cleaning- and disinfection<br />

agents. It is t<strong>here</strong>fore extremely important <strong>to</strong> target the bacteria before<br />

they ad<strong>here</strong> <strong>to</strong> the solid surface and develop in<strong>to</strong> a bi<strong>of</strong>ilm. One fac<strong>to</strong>r which may<br />

influence the adhesion <strong>of</strong> L. monocy<strong>to</strong>genes is the surface characteristics <strong>of</strong> processing<br />

equipment. In the present work, a method using fluorescence microscopy<br />

and a perfusion system will be developed <strong>to</strong> examine the initial adhesion <strong>of</strong> single<br />

L. monocy<strong>to</strong>genes cells <strong>to</strong> steel surfaces. Results will be presented showing the influence<br />

<strong>of</strong> flow direction on the adhesion <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> brushed stainless<br />

steel surfaces.<br />

Occurrence <strong>of</strong> Listeria monocy<strong>to</strong>genes in raw milk and dairy products<br />

in Kazerun, Iran<br />

Mehdi Mahmoodi, S. M. and Javanmardi, F.<br />

Islamic Azad University – Kazerun Branch, Iran<br />

The presence <strong>of</strong> Listeria monocy<strong>to</strong>genes was investigated in a <strong>to</strong>tal <strong>of</strong> 360 raw milk<br />

and dairy product samples including white cheese, yoghurt and Iranian yoghurt<br />

drink (Doogh) that were collected during five months from two traditional dairy<br />

manufacturer in Southern Iran. The prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in raw<br />

milk and white cheese samples <strong>of</strong> manufacturer A was found <strong>to</strong> be 1.7 % and 3.3 %<br />

respectively and <strong>of</strong> manufacturer B was found <strong>to</strong> be 3.3 % and 6.7 % respectively.<br />

No Listeria monocy<strong>to</strong>genes was isolated in any <strong>of</strong> the yoghurt and Doogh samples <strong>of</strong><br />

both manufacturer probably because <strong>of</strong> their low pH values.<br />

R E F E R E N C E<br />

D / P<br />

132<br />

R E F E R E N C E<br />

D / P<br />

133


R E F E R E N C E<br />

D / P<br />

134<br />

R E F E R E N C E<br />

D / P<br />

135<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Antilisterial mode <strong>of</strong> action <strong>of</strong> bacteriocin ST182Gu produced<br />

by Enterococcus casseliflavus isolated from guava<br />

Todorov, S. 1*, Destro, M. T. 1, Chiarini, E. B. 1, Vaz-Velho, M. 2 and Franco, B. D. G. M. 1<br />

1. University <strong>of</strong> Sao Paulo , Brazil<br />

2. ESTG, Viana do Castelo , Portugal<br />

This research is on the anti-Listeria bacteriocin ST182Gu and studies its mode <strong>of</strong> action.<br />

Bacteriocins <strong>of</strong> lactic acid bacteria (LAB) are ribosomally synthesized antimicrobial<br />

peptides. Their bactericidal mechanisms may include pore formation, degradation<br />

<strong>of</strong> cellular DNA, disruption through specific cleavage <strong>of</strong> 16S rDNA, and inhibition <strong>of</strong><br />

peptidoglycan synthesis. In our knowledge this is the first report <strong>of</strong> the isolation <strong>of</strong> bacteriocinogenic<br />

strain <strong>of</strong> Enterococcus casseliflavus from guava. Enterococcus casseliflavus<br />

ST182Gu, produces a pediocin-like bacteriocin (based on highly homology <strong>of</strong><br />

PCR product generated with the primers targeting pediocin PA-1 structural gene) with<br />

activity against several LAB, Listeria spp, and some other human and foodborne<br />

pathogens. No significant differences in growth and production <strong>of</strong> bacteriocin ST182Gu<br />

were observed when the strain E. casseliflavus ST182Gu was cultured for 24 h in MRS<br />

broth at 20 °C, 30 °C or 37 °C. At this 3 different cultivation temperatures activity was<br />

1.0 x10 9AU/ml against L. ivanovii subsp. ivanovii ATCC19119 and 2.6x10 4AU/ml against<br />

Enterococcus faecium ATCC19443. Low levels <strong>of</strong> bacteriocin ST182Gu activity (approximately<br />

1.3 x 10 4AU/ml against L. ivanovii subsp. ivanovii ATCC19119 were recorded<br />

after 3h <strong>of</strong> growth in MRS broth at 37°C and the maximal bacteriocin ST182Gu production<br />

(1.0 x 10 10AU/ml) was recorded in MRS after 27 – 33 h <strong>of</strong> growth. A decrease in activity<br />

was observed after 33 h <strong>of</strong> growth and it was reduced <strong>to</strong> 1.0x10 8AU/ml at 48 h <strong>of</strong><br />

incubation. Addition <strong>of</strong> bacteriocin ST182Gu <strong>to</strong> exponential or stationary phase cultures<br />

<strong>of</strong> L. ivanovii subsp. ivanovii ATCC19119 inhibited growth for 24 h. The effect <strong>of</strong><br />

bacteriocin ST182Gu on cells <strong>of</strong> L. ivanovii subsp. ivanovii ATCC19119 was visualised by<br />

AFM and indirectly recorded based on enzyme, protein and nucleotide material leakage.<br />

Bacteriocin ST182Gu is highly active against L. ivanovii subsp. ivanovii ATCC19119<br />

and exhibits a synergetic effect in the inhibition <strong>of</strong> this surrogate microorganism when<br />

applied <strong>to</strong>gether with subletal doses <strong>of</strong> cipr<strong>of</strong>loxacin.<br />

* Participation Supported by IUFoST.<br />

Control <strong>of</strong> Listeria monocy<strong>to</strong>genes in fresh goat cheese by bacteriocinogenic<br />

strain Lac<strong>to</strong>coccus lactis subsp. lactis DF4Mi or commercial nisin<br />

Nader Furtado, D., Todorov, S.*, Landgraf, M., Destro, M. T. and Franco, B. D. G. M.<br />

University <strong>of</strong> Sao Paulo , Brazil<br />

Bacteriocins are antimicrobial peptides produced by different groups <strong>of</strong> bacteria.<br />

Foods can be supplemented with bacteriocins, or by inoculation with the bacteriocinogenic<br />

strains, favouring the in situ production <strong>of</strong> the bacteriocin. The aim <strong>of</strong> this<br />

study was <strong>to</strong> evaluate the capability <strong>of</strong> the bacteriocinogenic strain Lac<strong>to</strong>coccus lactis<br />

subsp. lactis DF4Mi, isolated from goat milk, <strong>to</strong> control the growth Listeria monocy<strong>to</strong>genes<br />

in fresh goat cheese over 10 days <strong>of</strong> s<strong>to</strong>rage under refrigeration, and <strong>to</strong><br />

compare the effect that achieved in cheese added <strong>of</strong> commercial bacteriocin (nisin).<br />

Cheeses were prepared with 10 liters <strong>of</strong> pasteurized milk acidified with lactic acid<br />

(2.5 % v/v) and added <strong>of</strong> calcium chloride and commercial rennet, according <strong>to</strong> the<br />

cheese manufacturer practice. Three sets <strong>of</strong> cheeses were prepared: one with pasteurized<br />

milk with starter culture strain DF4Mi (10 6 UFC/ml), second with bacteriocin<br />

(-) strain <strong>of</strong> L. lactis subsp. lactis and other with addition <strong>of</strong> 12.5 mg/l <strong>of</strong> pure<br />

nisin (Fluka). L. monocy<strong>to</strong>genes 701 was added <strong>to</strong> cheeses during the manufacturing,<br />

in order <strong>to</strong> obtain 10 3 UFC/g. Appropriative controls such as cheese without any<br />

starter cultures, only L. monocy<strong>to</strong>genes 701, only strain DF4Mi or only bacteriocin (-<br />

) L. lactis subsp. lactis were prepared. Cheeses were s<strong>to</strong>red at 8 °C and submitted <strong>to</strong><br />

counts <strong>of</strong> lactic acid bacteria and L. monocy<strong>to</strong>genes every other day for 10 days period.<br />

Results indicate that L. monocy<strong>to</strong>genes decreased 3-log in the cheeses containing<br />

either bacteriocinogenic L. lactis subsp. lactis DF4Mi or bacteriocin (-) L. lactis<br />

subsp. lactis, showing that inhibition <strong>of</strong> L. monocy<strong>to</strong>genes might have occurred due<br />

<strong>to</strong> another fac<strong>to</strong>r than the production <strong>of</strong> bacteriocin. Analysis <strong>of</strong> the cheeses containing<br />

nisin demonstrated that incorporation <strong>of</strong> this bacteriocin at 12.5 mg/l could<br />

inhibit the growth <strong>of</strong> L. monocy<strong>to</strong>genes allowing a decrease <strong>of</strong> 2-log counts during<br />

s<strong>to</strong>rage. The present study showed that the use <strong>of</strong> purified bacteriocin in fresh goat<br />

cheese is more effective in the control <strong>of</strong> L. monocy<strong>to</strong>genes than the direct application<br />

<strong>of</strong> the bacteriocinogenic L. lactis subsp. lactis DF4Mi.<br />

* Participation Supported by IUFoST.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

High pressure processing ensures elimination <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in sliced dry cured ham<br />

S<strong>to</strong>llewerk, K., J<strong>of</strong>ré, A., Comaposada, J., Aymerich, T., Ferrini, G., Arnau, J. and Garriga, M.<br />

IRTA, Spain<br />

High pressure processing (HP) is a non-thermal treatment which inhibits or inactivates<br />

microorganisms and it is used <strong>to</strong> extend the shelf life <strong>of</strong> food products without<br />

altering their sensorial properties. Listeria monocy<strong>to</strong>genes can contaminate<br />

different types <strong>of</strong> meat products, such as dry cured ham, during slicing. The aim <strong>of</strong><br />

the present study was <strong>to</strong> evaluate the behavior <strong>of</strong> L. monocy<strong>to</strong>genes in three types<br />

<strong>of</strong> boned dry cured ham (H1, H2 and H3) produced following different processes.<br />

H1 was tumble salted and smoked; H2 was brine salted and fermented; H3 was<br />

brine salted, fermented and smoked. Afterwards, slices <strong>of</strong> ham were spiked with a<br />

mixture <strong>of</strong> 3 strains <strong>of</strong> L. monocy<strong>to</strong>genes (ca. 100 CFU/g). Drying was performed by<br />

the Quick Dry Slice process®. Pairs <strong>of</strong> slices were vacuum packed and half <strong>of</strong> them<br />

were submitted <strong>to</strong> a HP treatment <strong>of</strong> 600 MPa for 5 min. Packages were s<strong>to</strong>red<br />

under refrigeration for 112 days and the evolution <strong>of</strong> the pathogen was followed by<br />

plating in cromogenic agar and species-specific real time PCR. During s<strong>to</strong>rage <strong>of</strong><br />

non-pressurized hams the levels <strong>of</strong> L. monocy<strong>to</strong>genes progressively decreased in<br />

the three types <strong>of</strong> ham recording absence in 25 g in H3 after 112 days. Pressurization<br />

produced an immediate reduction <strong>of</strong> L. monocy<strong>to</strong>genes <strong>to</strong> levels below the<br />

plate detection limit (10 CFU/g) in all the hams. In H3 the pathogen was absent<br />

during the whole s<strong>to</strong>rage. After 112 days L. monocy<strong>to</strong>genes was not detected in any<br />

<strong>of</strong> the samples (absence in 25 g). The composition and processing <strong>of</strong> dry cured ham<br />

affected the survival <strong>of</strong> L. monocy<strong>to</strong>genes. Fermentation and smoking speed up the<br />

decrease <strong>of</strong> the pathogen, but only the application <strong>of</strong> a HP treatment enabled <strong>to</strong><br />

achieve a product free <strong>of</strong> L. monocy<strong>to</strong>genes during the whole s<strong>to</strong>rage period.<br />

Inhibition <strong>of</strong> Listeria monocy<strong>to</strong>genes by Lac<strong>to</strong>coccus spp. EU2241<br />

in tropical shrimp<br />

Abdoulaye Fall, P.<br />

Ifremer, France<br />

Shrimp is one <strong>of</strong> the most marketed seafood in the world. Due <strong>to</strong> its physicochemical<br />

parameters this product is very sensitive <strong>to</strong> microbial growth such as<br />

pathogens and spoiling microorganisms. Growth <strong>of</strong> these undesirable flora is <strong>of</strong>ten<br />

delayed by the use <strong>of</strong> classical techniques such as salting, low s<strong>to</strong>rage temperature,<br />

modification <strong>of</strong> atmosp<strong>here</strong> packaging or use <strong>of</strong> preservatives. However the<br />

human pathogen Listeria monocy<strong>to</strong>genes (Lm) that was evidenced in raw and ready<br />

<strong>to</strong> eat shrimp by some studies can <strong>of</strong>ten grow in those conditions and may reach<br />

the <strong>to</strong>lerated level <strong>of</strong> 100 Lm g -1. Biopreservation, an alternative technology for<br />

food preservation which consists in inoculating a product with selected bacteria <strong>to</strong><br />

eliminate or prevent growth <strong>of</strong> undesirable microorganisms, may be useful against<br />

Lm. In a previous study, a strain <strong>of</strong> Lac<strong>to</strong>coccus spp. EU2241 had been isolated<br />

from raw salmon and showed capacities <strong>to</strong> inhibit in model medium some<br />

pathogens isolated from seafood product. In this study the inhibition capacity <strong>of</strong><br />

Lac<strong>to</strong>coccus spp. EU2241 was tested in shrimp artificially contaminated with Lm.<br />

Four batches <strong>of</strong> cooked shrimp were prepared. 1: sterility control; 2: inoculation<br />

with Lac<strong>to</strong>coccus spp. (10 6 CFU g -1); 3: inoculation with Lm (10 3 CFU g -1); 4: co-inoculation<br />

<strong>of</strong> Lac<strong>to</strong>coccus and Lm. Shrimp were s<strong>to</strong>red at 8 °C for 31 days under<br />

modified atmosp<strong>here</strong> (CO 2 /N 2 50-50). When inoculated alone, Lm grew very well<br />

in shrimp reaching its maximum level, 10 9 CFU g -1, in 10 days. An immediate inhibition<br />

<strong>of</strong> Lm was observed in co-inoculated batch, reaching more than 4 logs at<br />

the end <strong>of</strong> s<strong>to</strong>rage with no sensory changes, confirming the interest <strong>of</strong> this bacterium<br />

for an application in biopreservation <strong>of</strong> seafood products against Lm. The<br />

use <strong>of</strong> Seafood Spoilage and Safety Predic<strong>to</strong>r s<strong>of</strong>tware <strong>to</strong> predict the growth <strong>of</strong> Lm<br />

in shrimp, showed that the lactic acid produced by Lac<strong>to</strong>coccus spp. (2500 ppm)<br />

and pH decrease from 6.60 <strong>to</strong> 5.91 could not <strong>to</strong>tally explain the inhibition observed<br />

in this study. A chemically defined liquid medium is currently developed <strong>to</strong> study<br />

the inhibition mechanisms.<br />

R E F E R E N C E<br />

D / P<br />

136<br />

R E F E R E N C E<br />

D / P<br />

137


R E F E R E N C E<br />

D / P<br />

138<br />

R E F E R E N C E<br />

D / P<br />

139<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Listeria monocy<strong>to</strong>genes and Listeria innocua in slaughter line <strong>of</strong> a swine<br />

meatpacking plant in Rio Grande do Sul State, Brazil<br />

Schittler, L. 1 and Padilha Silva, W. 2<br />

1. UDESC, Brazil<br />

2. UFPeL, Brazil<br />

Listeria monocy<strong>to</strong>genes is bacterium widely distributed in nature which presents<br />

high capacity <strong>to</strong> colonize surfaces and form bi<strong>of</strong>ilms in food processing plants. It is<br />

the causative agent <strong>of</strong> listeriosis which may cause a range <strong>of</strong> diseases from gastroenteritis<br />

<strong>to</strong> death, being the meat an important vehicle for transmission <strong>to</strong> humans.<br />

The aim <strong>of</strong> the present study was <strong>to</strong> investigate the presence and characterize<br />

L. monocy<strong>to</strong>genes and Listeria innocua in the slaughter line <strong>of</strong> a swine<br />

meatpacking plant, located in Missões Region, state <strong>of</strong> Rio Grande do Sul, Brazil.<br />

For this was evaluated water, cold chambers, carcasses, knives, faeces, hands and<br />

drains and the strains obtained were serotyped and submit <strong>to</strong> molecular typing<br />

through the analysis by random amplification <strong>of</strong> polymorphic DNA (RAPD) with<br />

UBC 155 e UBC 127 primers. The occurrence <strong>of</strong> L. monocy<strong>to</strong>genes was <strong>of</strong> 0.95 %<br />

and <strong>of</strong> L. innocua, 19.4 %, in 105 samples assayed, being the cold chambers the<br />

major point <strong>of</strong> isolation. Most <strong>of</strong> the L. innocua strains were not serotypable and,<br />

from those serotypable, the serovars 6a and 6b were prevalent. Only one L. monocy<strong>to</strong>genes<br />

strain was isolated, from a drain, however it belonged <strong>to</strong> serovar 4b, an<br />

important serovar in public health, because it is frequently involved in listeriosis in<br />

humans. RAPD <strong>of</strong> L. innocua strains produced 15 genetic combined pr<strong>of</strong>iles, whose<br />

index <strong>of</strong> discrimination was D=0.94, w<strong>here</strong>as the serotyping presented low discrimina<strong>to</strong>ry<br />

power (D=0.46). T<strong>here</strong> is contamination by L. monocy<strong>to</strong>genes e L. inoccua<br />

in the slaughter line <strong>of</strong> the swine meatpacking plant assayed, and the cross<br />

contamination in this plant is considered important, since some L. innocua strains<br />

persist in the local.<br />

Listeria monocy<strong>to</strong>genes in raw meat products marketed in the city<br />

<strong>of</strong> Sao Paulo, Brazil: Incidence and counts data for risk assessment<br />

Ris<strong>to</strong>ri, C. A. 1, Rowlands, R. E. G. 1, Martins, C. G. 1, Fávero, L. M. 2 and Franco, B. D. G. M. 3<br />

1. Food Microbiology Labora<strong>to</strong>ry, Adolfo Lutz Institute, Sao Paulo, SP, Brazil<br />

2. Food Surveillance, COVISA, Sao Paulo, SP, Brazil<br />

3. Department <strong>of</strong> Food and Experimental Nutrition, Faculty <strong>of</strong> Pharmaceutical Sciences, University<br />

<strong>of</strong> Sao Paulo, Sao Paulo, SP, Brazil<br />

Listeria monocy<strong>to</strong>genes is a ubiqui<strong>to</strong>us bacterium frequently found in meat products.<br />

Despite abundance <strong>of</strong> results on the prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in these<br />

foods in different countries, quantitative data on level <strong>of</strong> contamination required<br />

for appropriate risk assessments are scarce. This study aimed <strong>to</strong> evaluate the<br />

prevalence and numbers <strong>of</strong> Listeria monocy<strong>to</strong>genes in a variety <strong>of</strong> meat products<br />

available in the local market in the city <strong>of</strong> Sao Paulo, SP, Brazil. Between May 2008<br />

and July 2009, 552 samples <strong>of</strong> refrigerated raw meat products (138 beef sausages,<br />

138 pork sausages, 138 ground beef samples and 138 chicken leg samples) were<br />

purchased in 138 local supermarkets, selected after stratification by region and by<br />

random draw. Tests for presence and enumeration <strong>of</strong> L. monocy<strong>to</strong>genes were based<br />

on ISO 11290-1:1996/Amd.1:2004 and ISO 11290-2:1998 methods, respectively. L.<br />

monocy<strong>to</strong>genes was detected in 48.7% <strong>of</strong> the <strong>of</strong> meat products samples. The highest<br />

prevalence occurred in ground beef (59.4 %) followed by chicken legs (58 %), beef<br />

sausages (37.7 %), and pork sausages (39.8 %). In most samples (67.4 %), the level <strong>of</strong><br />

contamination was lower than 10 2 CFU/g. Counts ranged from < 10 <strong>to</strong> 5.6x10 2<br />

CFU/g in pork sausages, from < 10 <strong>to</strong> 1.9x10 2 CFU/g in beef sausages, from < 10 <strong>to</strong><br />

8.9x10 2 CFU/g in chicken legs, and from < 10 <strong>to</strong> 4.5x10 3 CFU/g in ground beef. Despite<br />

the low counts, data on prevalence <strong>of</strong> L. monocy<strong>to</strong>genes are relevant for estimating<br />

the risks <strong>of</strong> listeriosis associated <strong>to</strong> consumption <strong>of</strong> meat products in Sao<br />

Paulo, and for establishing science-based intervention strategies aimed at reducing<br />

these risks, specially for pregnant women and immunocompromised individuals.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Inhibi<strong>to</strong>n <strong>of</strong> Listeria monocy<strong>to</strong>genes by Carnobacterium<br />

maltaromaticum in combination with extract <strong>of</strong> Lippia sidoides Cham.<br />

in cold-smoked surubim fish broth<br />

Barbosa dos Reis, F., de Souza, V. M., Sousa Thomaz, M. R., Pin<strong>to</strong> Fernandes, L.,<br />

Pereira de Oliveira, W. and Pereira De Martinis, E. C.<br />

Faculdade de Ciências Farmacêuticas de Ribeirão Pre<strong>to</strong> - Universidade de São Paulo, Brazil<br />

Consumption <strong>of</strong> refrigerated cold-smoked fishes <strong>of</strong>fers risk <strong>of</strong> listeriosis, the severe<br />

foodborne infection caused by Listeria monocy<strong>to</strong>genes (Lm). In this work, <strong>to</strong><br />

improve the safety ready-<strong>to</strong>-eat fish, multiple hurdle experiments were carried<br />

out with cold-smoked surubim broth (CSSB) s<strong>to</strong>red at 5 °C for 5 weeks. To inhibit<br />

Lm, it was evaluated the application <strong>of</strong> a hidroalcoholic extract <strong>of</strong> the Brazilian<br />

folk plant Lippia sidoides (ELs) (10 µl/ml) in combination with Carnobacterium<br />

maltaromaticum (Cm) C2 and A9b + (bacteriocinogenic strains) and A9b - (nonbacteriocinogenic).<br />

CSSB was inoculated with 10 2 CFU <strong>of</strong> Lm/ml and/or 10 5 CFU<br />

<strong>of</strong> Cm/ml, as follow: 1) Lm; 2) Cm C2; 3) Cm A9b -; 4) Cm A9b +; 5) Lm plus ELs; 6)<br />

Cm C2 plus ELs; 7) Cm A9b - plus ELs; 8) Cm A9b + plus ELs; 9) Lm plus Cm C2; 10)<br />

Lm plus Cm A9b -; 11) Lm plus Cm A9b +; 12) Lm plus Cm C2 plus ELs; 13) Lm plus<br />

Cm A9b - plus ELs; 14) Lm plus Cm A9b + plus ELs. Bacterial populations and bacteriocin<br />

titer were determined after 24 h and once a week for up <strong>to</strong> 35 days. Control<br />

with Lm only, reached ca. 10 8 CFU/ml, but in co-culture with Cm, listerial<br />

population was significantly lower (ca. 10 4 CFU/ml with Cm A9b+ or Cm A9b- and<br />

< 10 2 CFU/ml with Cm C2). Bacteriocin production was observed only for Cm C2<br />

alone and Cm C2 plus Lm, which may explain the stronger inhibi<strong>to</strong>ry effect <strong>of</strong> Cm<br />

C2. ELs only did not inhibit Lm (ca. 10 8 CFU <strong>of</strong> Lm/ml) and its use in combination<br />

with any Cm strain did not present synergistic effect (ca. 10 4 CFU <strong>of</strong> Lm/ml). Bacteriocin<br />

production by Cm C2 was not detected in the presence <strong>of</strong> ELs, suggesting<br />

that ELs can adversely affect the protective culture and decrease the effectiveness<br />

<strong>of</strong> the biopreservation system studied.<br />

Inhibition <strong>of</strong> Listeria monocy<strong>to</strong>genes in cooked ham<br />

by virulent bacteriophages and protective cultures<br />

Holck, A., Schirmer, B. C. and Berg, J.<br />

N<strong>of</strong>ima Mat AS, Norway<br />

Listeria monocy<strong>to</strong>genes is found in a number <strong>of</strong> raw and ready-<strong>to</strong>-eat (RTE) products,<br />

poultry, seafood and dairy products. RTE products like cooked ham have little<br />

in<strong>here</strong>nt stability against the growth <strong>of</strong> L. monocy<strong>to</strong>genes. Several studies show<br />

reduction <strong>of</strong> L. monocy<strong>to</strong>genes in food when exposed <strong>to</strong> phages. Phage particles,<br />

however, become immobilised and are inactivated soon after addition <strong>to</strong> nonliquid<br />

foods and can thus not prevent later outgrowth <strong>of</strong> surviving Listeria. Here we have<br />

shown that protective cultures can be used successfully as an additional hurdle<br />

<strong>to</strong>gether with phages <strong>to</strong> reduce growth <strong>of</strong> L. monocy<strong>to</strong>genes. Sliced cooked ham<br />

was inoculated with cold adapted L. monocy<strong>to</strong>genes and exposed <strong>to</strong> bacteriophages.<br />

The hams were subsequently inoculated with Lac<strong>to</strong>bacillus sakei TH1 protective<br />

culture, vacuum packed and s<strong>to</strong>red at 10 °C. Addition <strong>of</strong> phages gave an immediate<br />

1-2 log reduction in L. monocy<strong>to</strong>genes. After 14 – 28 days <strong>of</strong> s<strong>to</strong>rage, an additional 2<br />

log reduction was observed in samples with phages and protective culture compared<br />

<strong>to</strong> samples with phages alone. The effect <strong>of</strong> the protective culture was evident<br />

both at 10 and 4 °C. Higher inoculation levels <strong>of</strong> protective culture gave a<br />

stronger inhibition <strong>of</strong> L. monocy<strong>to</strong>genes. The use <strong>of</strong> phages in combination with<br />

protective cultures is a general method that can potentially be applied <strong>to</strong> different<br />

products w<strong>here</strong> t<strong>here</strong> is a risk for growth <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

R E F E R E N C E<br />

D / P<br />

140<br />

R E F E R E N C E<br />

D / P<br />

141


R E F E R E N C E<br />

D / P<br />

142<br />

R E F E R E N C E<br />

D / P<br />

143<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Surface colonization by Listeria monocy<strong>to</strong>genes: role <strong>of</strong> the flagella<br />

in the bi<strong>of</strong>ilm formation process<br />

Desvaux, M., Briandet, R., Renier, S., Deschamps, J., NCaccia, N., Chafsey, I. and Hébraud, M.<br />

INRA, France<br />

The presence, contribution and/or requirement <strong>of</strong> exopolysaccharides in the<br />

bi<strong>of</strong>ilm formation process in Listeria monocy<strong>to</strong>genes have not been definitively ascertained.<br />

So far, the flagellum is the only cell surface fac<strong>to</strong>r demonstrated as involved<br />

in bi<strong>of</strong>ilm formation within this species, even so its exact contribution and<br />

role remain controversial. This prompted us <strong>to</strong> reinvestigate the role <strong>of</strong> the flagella<br />

in the course <strong>of</strong> bi<strong>of</strong>ilm formation at growth temperatures relevant <strong>to</strong> regulation<br />

<strong>of</strong> genetic expression in L. monocy<strong>to</strong>genes using both static and dynamic<br />

cultures conditions. Bi<strong>of</strong>ilm formation was assayed at early and late stages <strong>of</strong> the<br />

process using paramagnetic microbeads and crystal violet methods. Cell morphology<br />

and bi<strong>of</strong>ilm architecture were determined by confocal scanning laser microscopy<br />

using bacterial cells expressing fluorescent markers. Cells were grown<br />

statically in microtiter plates and in dynamic conditions using flow cells. As revealed<br />

by proteolytic treatments, extracellular/cell-surface proteins rather than<br />

exopolysaccharides are crucial determinants involved in listerial bi<strong>of</strong>ilm formation.<br />

While L. monocy<strong>to</strong>genes was enable <strong>to</strong> swim at 20 °C, no swarming could be<br />

observed. Kinetics <strong>of</strong> bi<strong>of</strong>ilm formation at 20 and 40 °C revealed that unflagellated<br />

mutant strain formed more bi<strong>of</strong>ilm than the wt except at the highest growth<br />

temperature tested. Besides crystal violet method, such results were confirmed<br />

by confocal microscopy in static and dynamic cultures conditions using fluorescently<br />

labelled listerial cells. Bi<strong>of</strong>ilm formation was for the first time investigated<br />

over a range <strong>of</strong> growth temperatures relevant <strong>to</strong> regulation <strong>of</strong> genetic expression in<br />

L. monocy<strong>to</strong>genes species. The use <strong>of</strong> complementary growth conditions, i.e. static<br />

and dynamic cultures, allowed providing a comprehensive picture <strong>of</strong> bi<strong>of</strong>ilm formation<br />

in this species. In those tested conditions it clearly appeared that flagella<br />

are not essential but rather detrimental <strong>to</strong> bi<strong>of</strong>ilm formation in L. monocy<strong>to</strong>genes.<br />

The role <strong>of</strong> sanitizers in controlling Listeria monocy<strong>to</strong>genes on stainless<br />

steel surfaces: Lessons learned from the 2008 Listeriosis outbreak<br />

Hébert, K., Farber, J. and Pagot<strong>to</strong>, F.<br />

Health Canada<br />

Proper maintenance and sanitization <strong>of</strong> food-contact and non-food-contact surfaces<br />

remains a high priority for industry in ensuring a microbiologically safe final<br />

product. Listeria monocy<strong>to</strong>genes is <strong>of</strong>ten present in fac<strong>to</strong>ry environments and<br />

needs <strong>to</strong> be controlled <strong>to</strong> avoid contamination <strong>of</strong> final product. The national outbreak<br />

<strong>of</strong> listeriosis caused by contaminated deli-meat raised questions about the<br />

ability <strong>of</strong> Listeria <strong>to</strong> survive on stainless steel surfaces. We evaluated the effectiveness<br />

<strong>of</strong> 11 sanitizers previously and currently used by Canadian industries<br />

against strains <strong>of</strong> L. monocy<strong>to</strong>genes. Eleven sanitizers were evaluated using the<br />

American Society for Testing and Materials standardized method (E2197) using<br />

stainless steel disk coupons, according <strong>to</strong> the recommendations on the labels or<br />

from the manufacturer. A soil load consisting <strong>of</strong> a microbial suspension, BSA,<br />

mucin and tryp<strong>to</strong>ne was used <strong>to</strong> mimic environmental conditions. Bacteria were<br />

inoculated at 10 7 CFU per coupon, dried and challenged against 50 µl <strong>of</strong> the suggested<br />

working concentration and exposure time <strong>of</strong> each sanitizer. The log reduction<br />

was calculated by direct plating. Ten sanitizers met the performance criteria<br />

for L. monocy<strong>to</strong>genes, achieving a minimal 6-log reduction. T<strong>here</strong> was no significant<br />

difference observed amongst the different active ingredient formulations. A<br />

single, alchohol-based hand sanitizer used by the industry did not satisfy the criteria<br />

when tested on stainless steel coupons. This study demonstrated that sanitizers<br />

are effective in controlling Listeria on stainless steel surfaces when proper<br />

ad<strong>here</strong>nces <strong>to</strong> pro<strong>to</strong>cols were maintained. Misuse <strong>of</strong> sanitizers can allow the organism<br />

<strong>to</strong> survive and possibly make its way in<strong>to</strong> the final food product.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Listeriophage ecology and diversity on dairy farms<br />

Vongkamjan, K.*, Moreno Switt, A., den Bakker, H. C., Fortes, E. D., and Wiedmann, M.<br />

Department <strong>of</strong> Food Science, Cornell University, Ithaca, NY, USA<br />

Listeria monocy<strong>to</strong>genes is common in dairy farm environments and contaminated<br />

silage appears <strong>to</strong> be the major source <strong>of</strong> L. monocy<strong>to</strong>genes infections in farm ruminants.<br />

Improved understanding <strong>of</strong> the ecology and diversity <strong>of</strong> phages infecting<br />

L. monocy<strong>to</strong>genes (“listeriophages”) in food associated environments is critical<br />

<strong>to</strong> evaluate the impact, effects, and potential <strong>of</strong> phage-based pathogen control<br />

strategies. We collected silage samples over the course <strong>of</strong> one year on two dairy<br />

farms <strong>to</strong> perform screening for L. monocy<strong>to</strong>genes and listeriophages, using standard<br />

methods and representative host strains. L. monocy<strong>to</strong>genes isolates were characterized<br />

by molecular methods (e.g., sequencing <strong>of</strong> sigB) and subsequently classified<br />

in<strong>to</strong> lineages I and II. Listeriophages were characterized for host range using<br />

a set <strong>of</strong> reference strains that represented 13 L. monocy<strong>to</strong>genes serotypes. Furthermore,<br />

phage genome size was determined by the Pulsed Field Gel Electrophoresis.<br />

Out <strong>of</strong> 134 silage samples tested (farm 1, n = 81; farm 2, n = 53), six<br />

samples were positive for L. monocy<strong>to</strong>genes; 43/81 samples from farm 1 and 40/53<br />

samples from farm 2 were positive for listeriophages. Overall 34/134 silage samples<br />

contained >100 PFU/g silage with samples containing up <strong>to</strong> 2x10 4 PFU/g. For the<br />

majority <strong>of</strong> phage positive-samples, phages were detected on serotype 4a host<br />

strain. Phage isolates represented considerable host range diversity (12 distinct<br />

lysis groups). Most phages did not lyse lineage I strains, but lysed most lineage II<br />

strains and showed consistent and strong lysis patterns with all lineage III strains.<br />

Phage genome size ranged from 25 <strong>to</strong> 120 kb. Our data show considerable listeriophage<br />

diversity in dairy farm environments, suggesting that phages play an important<br />

role in the ecology <strong>of</strong> L. monocy<strong>to</strong>genes on dairy farms.<br />

* Participation Supported by IUFoST.<br />

Evaluation <strong>of</strong> curative and preventive decontamination treatments<br />

on Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms with a new screening system<br />

Quinon, E. 1, Chamot, S. 1, Groelly, J. 2, Chavant, P. 2, Bernardi, T. 2, Desvaux, M. 1<br />

and Hebraud, M. 1<br />

1. INRA, France<br />

2. BioFilm Control, France<br />

Food safety requires not only interventions on the product but also on all surfaces<br />

<strong>of</strong> the workshops from the receipt <strong>of</strong> raw materials until the processing and packaging<br />

<strong>of</strong> food products. All surfaces can be contaminated by microorganisms able <strong>to</strong><br />

ad<strong>here</strong>, <strong>to</strong> form bi<strong>of</strong>ilms and <strong>to</strong> disseminate all along the chain. The literature<br />

shows that bacterial cells in bi<strong>of</strong>ilm, compared with their plank<strong>to</strong>nic counterpart,<br />

present a greater resistance <strong>to</strong> hostile environments and particularly <strong>to</strong> cleaning<br />

disinfection procedures. The efficacy <strong>of</strong> cleaning disinfection products currently<br />

used for surface decontamination is generally evaluated on plank<strong>to</strong>nic cells, which<br />

does not predict their efficacy on bacteria in bi<strong>of</strong>ilm. We have implemented a<br />

screening system, initially developed by the Bi<strong>of</strong>ilm Control society <strong>to</strong> evaluate<br />

bacterial ability <strong>to</strong> form a bi<strong>of</strong>ilm on an abiotic surface (BioFilm Ring Test â), <strong>to</strong> test<br />

the efficacy <strong>of</strong> 4 commercial decontamination solutions <strong>to</strong> detach bi<strong>of</strong>ilm from the<br />

surface and <strong>to</strong> affect the viability <strong>of</strong> bacterial cells. Three temperatures (10 °C, 20 °C<br />

and 37 °C) and three Listeria monocy<strong>to</strong>genes strains were used for the tests. The effect<br />

<strong>of</strong> preventive decontamination treatment <strong>of</strong> the surface on cell adhesion and<br />

bi<strong>of</strong>ilm formation was also assessed. Two out <strong>of</strong> the 4 products, containing quaternary<br />

ammonium compound (QAC) as biocide, dramatically affected the viability<br />

<strong>of</strong> sessile cells whatever the temperature while only one out <strong>of</strong> the two was efficient<br />

<strong>to</strong> detach ad<strong>here</strong>d cells. The formulation <strong>of</strong> this last product also contains a<br />

polyenzymatic cocktail. The two other solutions remained inefficient on bi<strong>of</strong>ilms.<br />

The QAC + polyenzymatic cocktail was also the only product efficient <strong>to</strong> prevent or<br />

delay the bi<strong>of</strong>ilm formation after a preventive treatment <strong>of</strong> the surface. The<br />

BioFilm Ring Test â can be used <strong>to</strong> sreen the efficacy <strong>of</strong> biocides on bacterial bi<strong>of</strong>ilms<br />

in order <strong>to</strong> improve the microbiological status <strong>of</strong> the food processing plants.<br />

R E F E R E N C E<br />

D / P<br />

144<br />

R E F E R E N C E<br />

D / P<br />

145


R E F E R E N C E<br />

D / P<br />

146<br />

R E F E R E N C E<br />

D / P<br />

147<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Pulsed Field Gel Electrophoresis, conventional and molecular<br />

serotyping on Listeria monocy<strong>to</strong>genes: An European Pr<strong>of</strong>iciency<br />

Testing Inter-labora<strong>to</strong>ry Trial<br />

Félix, B., Roussel, S., Tam Dao, T., Asséré, A., Lombard, B. and Brisabois, A.<br />

Agence française de sécurité sanitaire des aliments (Afssa), Labora<strong>to</strong>ire d’Etudes et de Recherches sur<br />

la Qualité des Aliments et sur les Procédés agro-alimentaires (Lerqap), Maisons-Alfort, France<br />

In Europe, the incidence <strong>of</strong> listeriosis caused by Listeria monocy<strong>to</strong>genes is relatively<br />

low, however an increase in cases, which differ depending <strong>of</strong> the countries,<br />

has been observed since 2000. Different methods are available for first-level characterization<br />

in epidemiological surveillance, in particular, conventional and molecular<br />

serotyping. Regarding fine characterization techniques, macro-restriction<br />

analysis by pulsed-field gel electrophoresis (PFGE) has been described as the most<br />

discrimina<strong>to</strong>ry methods for L. monocy<strong>to</strong>genes subtyping. This method is widely<br />

used by the European National Reference Labora<strong>to</strong>ries (NRLs) according <strong>to</strong> a<br />

2009 EFSA survey. AFSSA’s Labora<strong>to</strong>ry for Study and Research on Food Quality<br />

and Processing was appointed as Community Reference Labora<strong>to</strong>ry (CRL) for Listeria<br />

monocy<strong>to</strong>genes in 2006. Recently, the CRL organized a pr<strong>of</strong>iciency test in<br />

order <strong>to</strong> evaluate the ability <strong>of</strong> the NRLs <strong>to</strong> perform these three typing methods.<br />

Labora<strong>to</strong>ries were free <strong>to</strong> choose either one, two or all three methods. The same<br />

panel <strong>of</strong> six strains was used for all three methods and was sent <strong>to</strong> the 17 participating<br />

NRLs. This panel included serotypes 1/2a, 4b, 1/2c, 1/2b and 3a strains. Each<br />

labora<strong>to</strong>ry was free <strong>to</strong> choose the pro<strong>to</strong>cols used. All participants followed the<br />

PFGE PulseNet Europe standardized pro<strong>to</strong>col. PFGE migration parameters were<br />

set by the CRL, pattern interpretations were performed following the CRL’s interpretation<br />

criterion. For conventional serotyping and molecular serotyping, 86 %<br />

and 93.5 % <strong>of</strong> the serotypes obtained were in agreement with the CRL data. Inconsistencies<br />

were mostly explained by differences observed in the use <strong>of</strong> reagents.<br />

For PFGE, 83 % and 81 % <strong>of</strong> pr<strong>of</strong>iles for AscI-PFGE and for ApaI-PFGE was undistinguishable<br />

<strong>of</strong> the expected pr<strong>of</strong>iles The inconsistencies observed were caused<br />

by slight standardization defaults or in few cases by a fault in the extraction. This<br />

PT trial was a valuable opportunity <strong>to</strong> improve the typing ability <strong>of</strong> NRLs and will<br />

allow PFGE pattern exchanges.<br />

Detection <strong>of</strong> Listeria spp. in raw and pasteurized liquid egg-products<br />

and in the egg-breaking plants environment<br />

Rivoal, K., Fablet, A., Chemaly, M., Salvat, G. and Protais, J.<br />

AFSSA, France<br />

Listeria monocy<strong>to</strong>genes has been recognized as a human pathogen for decades and<br />

is etablished as an important food-borne pathogen. T<strong>here</strong> is a lack <strong>of</strong> information<br />

about eggs and egg-products contamination by Listeria spp. The aim <strong>of</strong> this study<br />

was <strong>to</strong> detect Listeria spp. in liquid egg-products (whole, yolk and white with or<br />

without salt and/or sugar) collected in 5 egg-breaking plants in north western<br />

France, during one year. To study the seasonal variation, each egg-breaking plant<br />

has been visited during each season. For a raw egg-product sample, a pasteurized<br />

sample has been tested and the egg shells used for theses egg-products have been<br />

swabed <strong>to</strong> detect Listeria spp. The environment <strong>of</strong> plants was also sampled with<br />

swabs during the production <strong>of</strong> the egg-products tested. At the present time, the<br />

project is not achieved. Three seasons have been studied and the last one is now<br />

going on. The environment <strong>of</strong> the egg-breaking plants seems <strong>to</strong> be very contaminated<br />

by Listeria spp. when the egg-products and their shells were slightly contaminated<br />

by the bacteria. Moreover, Listeria monocy<strong>to</strong>genes has been very rarely<br />

isolated in the egg-products as well as in the environment. At the end <strong>of</strong> the year,<br />

the study will be finished and the results could be analysed.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Bi<strong>of</strong>ilm formation by Listeria monocy<strong>to</strong>genes isolates under conditions<br />

that mimic food and digestive tract<br />

Kretli Winkelströter, L. 1, Oliveira, M. A. 2 and De Martinis, E. C. 1<br />

1. Universidade de São Paulo, Brazil<br />

2. Institu<strong>to</strong> Adolfo Lutz, Brazil<br />

L. monocy<strong>to</strong>genes is ubiqui<strong>to</strong>us; able <strong>to</strong> <strong>to</strong>lerate many adverse conditions found in<br />

foods and can persist in processing plants due <strong>to</strong> formation <strong>of</strong> bi<strong>of</strong>ilms. L. monocy<strong>to</strong>genes<br />

can cause severe foodborne infections in immunocompromised people<br />

and in pregnant women. In this work, L. monocy<strong>to</strong>genes isolates were screened for<br />

ability <strong>to</strong> form bi<strong>of</strong>ilm under diverse conditions. Overnight cultures <strong>of</strong> 51 isolates<br />

<strong>of</strong> L. monocy<strong>to</strong>genes from food, clinical and environmental sources were inoculated<br />

(1:20, v/v) in Brain Heart Infusion Broth (BHI) and under the following conditions:<br />

a) BHI plus 2 % sucrose; b) BHI plus 5 % NaCl; c) BHI plus 0.3 % Oxgall; d)<br />

BHI pH 2 (adjusted with HCl) and e) BHI plus 50 % (v/v) <strong>of</strong> extract from bacteriocin-producing<br />

cultures <strong>of</strong> Lac<strong>to</strong>bacillus sakei 1, Leuconos<strong>to</strong>c mesenteroides 11A or<br />

Enterococcus faecium 130. Aliquots <strong>of</strong> inoculated broths (200µl) were transferred<br />

<strong>to</strong> wells <strong>of</strong> a polystyrene 96-wells microtiter plate and incubated at 37 °C/24h.<br />

Wells were washed 3 times with Phosphate Buffered Saline (PBS) and each well<br />

was stained with 200µL <strong>of</strong> 1 % (v/v) crystal violet aqueous solution for 15 min.<br />

Plates were washed 3 times with PBS and wells were destained with 200µL <strong>of</strong> 95 %<br />

ethanol. Concentration <strong>of</strong> crystal violet present in the destaining solution was<br />

measured at 595 nm (CV-OD 595 ) and directly correlated with bi<strong>of</strong>ilm formation. L.<br />

monocy<strong>to</strong>genes ATCC 19115 was used as parameter <strong>of</strong> comparison for bi<strong>of</strong>ilm formation.<br />

L. monocy<strong>to</strong>genes ATCC 19115 presented CV-OD 595 in the interval <strong>of</strong> 0.69-<br />

0.87 under all conditions tested. CV-OD 595 >1.1 was observed for approximately<br />

9.8 %, 1.9 %, 7.8 % and 19.6 % for L. monocy<strong>to</strong>genes isolates incubated respectively in<br />

BHI only and under presence <strong>of</strong> sucrose, NaCl and Oxgall, indicating a higher ability<br />

<strong>to</strong> form bi<strong>of</strong>ilm. In BHI pH 2 and BHI plus 50 % (v/v) <strong>of</strong> extract from bacteriocin-producing<br />

cultures, L. monocy<strong>to</strong>genes isolates did not show higher differences<br />

in bi<strong>of</strong>ilm formation in comparison with L. monocy<strong>to</strong>genes ATCC 19115 (0.70-0.81<br />

CV-OD 595 ). L. monocy<strong>to</strong>genes isolates showed different abilities <strong>to</strong> form bi<strong>of</strong>ilm<br />

under different conditions probably due <strong>to</strong> modulation <strong>of</strong> gene expression, which<br />

will be further studied.<br />

Bi<strong>of</strong>ilm formation <strong>of</strong> Listeria monocy<strong>to</strong>genes EGDe depends<br />

on temperature and nutrient availability<br />

Auchter, M.*, Endres, J., Waidmann, M. S. and Riedel, C. U.<br />

Institute <strong>of</strong> Microbiology and Biotechnology, University <strong>of</strong> Ulm, Ulm, Germany<br />

Bioluminescent in vivo imaging revealed that in mice, Listeria monocy<strong>to</strong>genes colonizes<br />

the gall bladder prior <strong>to</strong> systemic spread. Moreover, bacterial cells <strong>of</strong> L.<br />

monocy<strong>to</strong>genes isolated from the gall bladder display chaining morphology, a phenotype<br />

associated with bi<strong>of</strong>ilm grown cells. Thus, bi<strong>of</strong>ilm formation in the gall<br />

bladder could be a strategy <strong>of</strong> this important pathogen <strong>to</strong> protect, establish and<br />

multiply itself at a site inaccessible <strong>to</strong> the components <strong>of</strong> the immune system prior<br />

<strong>to</strong> causing systemic infections. To investigate the mechanisms <strong>of</strong> bi<strong>of</strong>ilm formation,<br />

we tested L. monocy<strong>to</strong>genes EGDe and an isogenic agrD deletion mutant for<br />

their ability <strong>to</strong> form bi<strong>of</strong>ilms under different conditions. Bi<strong>of</strong>ilm formation was<br />

tested in a standard microtiter plate assay at different time points during bi<strong>of</strong>ilm<br />

growth <strong>of</strong> EGDe, varying temperatures (20, 30 and 37 °C) and in full-strength BHI<br />

(rich broth) as well as in 10-fold diluted BHI medium (nutrient limited). Bi<strong>of</strong>ilm<br />

formation at 37 °C was poor in rich broth and could be significant enhanced when<br />

cells were grown under nutrient limitation. Deletion <strong>of</strong> the agrD gene resulted in<br />

defective bi<strong>of</strong>ilms in 10-fold diluted, as well as in full-strength BHI. The bi<strong>of</strong>ilm<br />

formation <strong>of</strong> the ∆agrD mutant could also be res<strong>to</strong>red <strong>to</strong> wild type levels if reconstituted<br />

cell-free supernatant (BHI added <strong>to</strong> spent EGDe supernatant, pH adjusted<br />

and filter sterilized) <strong>of</strong> EGDe wild type cells grown in full-strength BHI was added<br />

<strong>to</strong> the ∆agrD strain. Additionally, DDAO staining and DNAse treatments revealed<br />

a potential role for extracellular DNA in the initial stages <strong>of</strong> bi<strong>of</strong>ilm formation.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

D / P<br />

148<br />

R E F E R E N C E<br />

D / P<br />

149


R E F E R E N C E<br />

D / P<br />

150<br />

R E F E R E N C E<br />

D / P<br />

151<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Antimicrobial activity <strong>of</strong> lac<strong>to</strong>coccal and enterococcal strains<br />

isolated from artisanal products from North West <strong>of</strong> Italy <strong>to</strong>ward<br />

Listeria monocy<strong>to</strong>genes<br />

Dal Bello, B., Rantsiou, K., Ambrosoli, R., Zeppa, G. and Cocolin, L.<br />

Department <strong>of</strong> Exploitation and Protection <strong>of</strong> the Agricultural and Forestry Resources, University <strong>of</strong><br />

Turin, Italy<br />

Listeria monocy<strong>to</strong>genes is a foodborne pathogenic Gram-positive bacterium that is<br />

widely distributed in soil, sewage, fresh water sediments and effluents, and is frequently<br />

carried in the intestinal tract <strong>of</strong> animals and humans. Biocontrol <strong>of</strong> L.<br />

monocy<strong>to</strong>genes by bacteriocin-producing lactic acid bacteria or by bacteriocin extracts<br />

has attracted great attention in recent years and new preservation strategies<br />

<strong>to</strong> control growth <strong>of</strong> L. monocy<strong>to</strong>genes have been developed, including the application<br />

<strong>of</strong> the bacteriocin nisin (Food and Drug Administration, 1988). In this work,<br />

we have investigated the potential role as bioprotection agents <strong>of</strong> au<strong>to</strong>chthonous<br />

lac<strong>to</strong>coccal and enterococcal strains isolated from fresh and fermented artisanal<br />

products (cheese and meat) <strong>of</strong> Piedmont region (North West <strong>of</strong> Italy) determining<br />

their antimicrobial spectrum <strong>of</strong> activity <strong>to</strong>wards different foodborne spoilage and<br />

pathogenic microorganisms. Bacteriocin-producing strains were identified by molecular<br />

methods and genetic determinants encoding the antimicrobial proteins<br />

were targeted by PCR. Thirty-nine strains <strong>of</strong> Lac<strong>to</strong>coccus lactis exhibited inhibition<br />

<strong>to</strong>wards L. monocy<strong>to</strong>genes NCTC 10527 and most <strong>of</strong> them (26 strains) showed<br />

the presence <strong>of</strong> the genes responsible for the nisins A and Z production. Regarding<br />

Enterococcus spp., 31 strains showed inhibition activity <strong>to</strong>wards L. monocy<strong>to</strong>genes<br />

NCTC 10527 and the presence <strong>of</strong> the genes responsible for the enterocins A and P<br />

production was determined. It is interesting <strong>to</strong> underline that for some strains it<br />

was not possible <strong>to</strong> identify any known bacteriocins. In this study a high incidence<br />

<strong>of</strong> bacteriocin producing strains was observed. In the future, the possible use <strong>of</strong><br />

these active strains could be a new way <strong>to</strong> ensure safety <strong>of</strong> foods.<br />

Plant-based strategies for Listeria monocy<strong>to</strong>genes control in foods<br />

Paparella, A., Serio, A., Chaves-Lopez, C. and Di Pasquale, F.<br />

University <strong>of</strong> Teramo, Italy<br />

Essential oils (EOs) and plant extracts are considered a natural and effective alternative<br />

<strong>to</strong> chemical preservatives. Considering the complex chemical composition<br />

<strong>of</strong> EOs, their antimicrobial action is not likely attributable <strong>to</strong> one single mechanism,<br />

although bacterial cy<strong>to</strong>plasmic membrane seems <strong>to</strong> be a specific target.<br />

The authors evaluated the antimicrobial activity <strong>of</strong> thyme, oregano and cinnamon<br />

EOs against Listeria monocy<strong>to</strong>genes strains, isolated from the smoked salmon industry<br />

and from meat products; these strains showed variable levels <strong>of</strong> antimicrobial<br />

resistance, belonged <strong>to</strong> different serotypes and had heterogeneous molecular<br />

pr<strong>of</strong>iles. Thyme and oregano EOs were particularly effective, with Minimal Inhibi<strong>to</strong>ry<br />

Concentrations (MICs) correlated with strains biodiversity. Au<strong>to</strong>matic<br />

turbidometry was applied <strong>to</strong> evaluate the effects <strong>of</strong> EOs on cell physiology; the results<br />

documented a lag phase extension and a decrease <strong>of</strong> maximum growth rate<br />

and maximum growth value. The antibacterial effect was evident even after a short<br />

contact time with the cells. Using flow cy<strong>to</strong>metry combined with fluorescent techniques,<br />

the authors demonstrated that membrane disruption is the primary inactivation<br />

mechanism <strong>of</strong> thyme and oregano oils. A different mechanism might be involved<br />

in cinnamon EO, acting on cells enzymatic activity. Further studies, carried<br />

out by means <strong>of</strong> EPR (Electronic Paramagnetic Resonance), highlighted the effect<br />

<strong>of</strong> oregano EO on membrane fluidity and order, variable with increasing EO concentrations.<br />

These findings prove that food biopreservation with EOs involves various<br />

mechanisms <strong>of</strong> action and may be considered a safe and effective measure <strong>to</strong><br />

control Listeria monocy<strong>to</strong>genes growth, although with different effects depending<br />

on strains biodiversity.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Scientific studies for survival <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

in dairy products and its application in practice<br />

Cabanova, L., Škun<strong>to</strong>va, O. and Kantikova, M.<br />

State Veterinary and Food Institute, Dolny Kubin, Slovakia<br />

European reference labora<strong>to</strong>ry for Listeria monocy<strong>to</strong>genes in Paris has prepared<br />

<strong>to</strong>gether with some National reference labora<strong>to</strong>ries from the European Member<br />

States the scientific study for Listeria <strong>to</strong> prove the producer that his products are<br />

safe till the end <strong>of</strong> the shelf- life. In our study we have developed studies for the<br />

producer <strong>of</strong> two different dairy products after several positive findings <strong>of</strong> Listeria<br />

in the final products. As a first product the raw sheep cheese has been selected<br />

and as a second one the steamed cheese products have been analyzed. In the first<br />

stage the physico-chemical characteristics has been determined (a w , pH) and than<br />

the subsamples has been prepared and artificially contaminated with a mixture<br />

<strong>of</strong> two reference Listeria strains and third isolated strain from the previous positive<br />

sample. After inoculation treated samples has been s<strong>to</strong>red at 5.8–6.2 °C incuba<strong>to</strong>r<br />

till the end <strong>of</strong> the shelf-life and during this period 2-3 times has been analyzed<br />

(a w , pH, Listeria detection and enumeration has been determined). At the<br />

end <strong>of</strong> the shelf – life the growth potential has been calculated and the Listeria<br />

numbers at the beginning and at the end <strong>of</strong> the shelf- life has been given.<br />

The effect <strong>of</strong> chilling temperatures on the virulence<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes isolates with different origins<br />

Neves, E. M. 1,2, Silva, A. C. 1*, Louro, P. 3,4, Ferreira-Dias, S. 4 and Bri<strong>to</strong>, L. 1<br />

1. CBAA/ Departamen<strong>to</strong> de Botânica e Engenharia Biológica, Institu<strong>to</strong> Superior de Agronomia,<br />

Technical University <strong>of</strong> Lisbon, Portugal<br />

2. Institu<strong>to</strong> Superior de Estudos Interculturais e Transdisciplinares (ISEIT), Campus Universitário<br />

de Almada, Institu<strong>to</strong> Piaget, Portugal<br />

3. Institu<strong>to</strong> Nacional dos Recursos Biológicos, IP, Portugal<br />

4. CEER/ Departamen<strong>to</strong> de Agro-Indústrias e Agronomia Tropical, Institu<strong>to</strong> Superior de Agronomia,<br />

Technical University <strong>of</strong> Lisbon, Portugal<br />

Although the chilling and freezing <strong>of</strong> food products are increasingly used by the<br />

food industry, t<strong>here</strong> are few studies on the effect <strong>of</strong> these temperatures on the virulence<br />

<strong>of</strong> Listeria monocy<strong>to</strong>genes. The influence <strong>of</strong> 1, 7 and 30 days <strong>of</strong> s<strong>to</strong>rage at<br />

chilling (7 °C) and freezing temperatures (-20 °C and -76 °C) on the virulence <strong>of</strong><br />

19 L. monocy<strong>to</strong>genes isolates was evaluated in this study. The isolates selected according<br />

<strong>to</strong> their genetic diversity and different initial levels <strong>of</strong> virulence were from<br />

frozen (n = 5), refrigerated (n = 3), refrigerated ready-<strong>to</strong>-eat (n = 5) and ready-<strong>to</strong>eat<br />

foods (n = 1), from humans (n = 2), from dairy environment (n = 1) and reference<br />

strains (n = 2). In the case <strong>of</strong> the chilling assays, the cells were kept in physiological<br />

buffer and culture medium was added just before infection. After the<br />

respective s<strong>to</strong>rage, the isolates were used <strong>to</strong> infect HT-29 cell monolayers in a<br />

plaque forming assay (pfa), immediately or after thawing the bacterial suspensions<br />

at room temperature. The pathogenic potential <strong>of</strong> the strains was expressed<br />

as the mean log <strong>of</strong> the number <strong>of</strong> plaques formed (log pfa), and one-way or multifac<strong>to</strong>rial<br />

ANOVA <strong>of</strong> the counting data was carried out. The results showed that the<br />

refrigeration temperature (7 °C) acted more significantly (P < 0.05) on the decreasing<br />

<strong>of</strong> the virulence potential <strong>of</strong> all isolates, although time didn´t show a significant<br />

effect on this loss <strong>of</strong> virulence. In relation <strong>to</strong> freezing temperatures (-<br />

20 °C and -76 °C), the decrease in virulence was not as significant, although the<br />

temperature <strong>of</strong> -76 °C tended <strong>to</strong> be more conservative <strong>of</strong> the initial virulence. Better<br />

elucidation <strong>of</strong> which refrigerated food matrices allow growth <strong>of</strong> the pathogen<br />

may help <strong>to</strong> evaluate the real risk <strong>of</strong> the presence <strong>of</strong> Listeria in food.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

D / P<br />

152<br />

R E F E R E N C E<br />

D / P<br />

153


R E F E R E N C E<br />

D / P<br />

154<br />

R E F E R E N C E<br />

D / P<br />

155<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

How <strong>to</strong> improve a sampling plan in order <strong>to</strong> better assess<br />

L. monocy<strong>to</strong>genes contamination on diced bacon at the plant<br />

Bergis, H. 1, Commeau, N. 1, Zuliani, V. 2, Cornu, M. 1, Beaufort, A. 1 and Garry, P. 2<br />

1. AFSSA, France<br />

2. IFIP, Institut de la filière porcine, France<br />

Food Business Opera<strong>to</strong>rs (FBO’s) are legally responsible for the safety <strong>of</strong> their food<br />

products (Regulation CE 2073/2005). The regulation does not specify the frequency<br />

<strong>of</strong> sampling/ testing and it is up <strong>to</strong> the FBO’s <strong>to</strong> decide the appropriate<br />

level <strong>of</strong> sampling/testing <strong>to</strong> help validate and verify their food safety management<br />

plans. In this scope, sampling plans are useful <strong>to</strong>ols <strong>to</strong> control and <strong>to</strong> assess the<br />

contamination <strong>of</strong> a given foodborne pathogen at the plant and also <strong>to</strong> help the FBO<br />

<strong>to</strong> take decisions. In the present study, a production <strong>of</strong> diced bacon in a French<br />

plant was moni<strong>to</strong>red <strong>to</strong> assess the contamination <strong>of</strong> this product by Listeria monocy<strong>to</strong>genes.<br />

The steps <strong>of</strong> the process are: tumbling, steaming, dicing, and packaging.The<br />

aim <strong>of</strong> this study was <strong>to</strong> try, through the elaborated sampling plan, (i) <strong>to</strong> establish<br />

a link between the prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in the pork breasts and<br />

in diced bacon; (ii) <strong>to</strong> determine whether the levels <strong>of</strong> contamination <strong>of</strong> the pork<br />

breasts and in the diced bacon are related; (iii) <strong>to</strong> estimate the incidence <strong>of</strong> the<br />

tumbling step and <strong>of</strong> the dicing step on the final product contamination.The food<br />

process was studied from the raw material (pork breasts) <strong>to</strong> the final product<br />

(diced bacon) and the steps possibly involved in the contamination identified. An<br />

experimental stratified random sampling plan was constructed in two steps <strong>of</strong> the<br />

process: after tumbling (a important step), and after packaging. 106 sampled tumbling<br />

breast and 86 packages <strong>of</strong> diced bacon were analysed. Detection and enumeration<br />

<strong>of</strong> L. monocy<strong>to</strong>genes were performed on 100cm2 <strong>of</strong> the pork breast and on<br />

100g <strong>of</strong> diced bacon. The lactic acid bacteria (LAB) were also enumerated. From<br />

the obtained data on contamination (prevalence and level <strong>of</strong> contamination <strong>of</strong> L.<br />

monocy<strong>to</strong>genes and the level <strong>of</strong> LAB), it appears that the tumbling step has an homogenisation<br />

effect on the contamination and sampling after this step could be a<br />

good indica<strong>to</strong>r <strong>of</strong> the presence <strong>of</strong> L. monocy<strong>to</strong>genes in diced bacon. It has been<br />

shown that over a certain level <strong>of</strong> contamination <strong>of</strong> the breast, the diced bacon<br />

from the same batch are also contaminated with L. monocy<strong>to</strong>genes.<br />

Contamination <strong>of</strong> Listeria monocy<strong>to</strong>genes in a cold-smoked<br />

pork processing plant using brining injections<br />

Berzins, A. 1, Silins, I. 1 and Korkeala, H. 2<br />

1. Faculty <strong>of</strong> Veterinary Medicine, Latvia University <strong>of</strong> Agriculture; University <strong>of</strong> Helsinki, Latvia<br />

2. University <strong>of</strong> Helsinki, Finland<br />

Contamination <strong>of</strong> L. monocy<strong>to</strong>genes was studied in a cold-smoked pork processing<br />

plant using brining injections. Overall prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in coldsmoked<br />

pork products during a 5-year period was 26 %. Environmental sampling<br />

combined with PFGE subtyping and serotyping was applied <strong>to</strong> investigate the genetic<br />

diversity <strong>of</strong> L. monocy<strong>to</strong>genes in the brining facilities and alongside premises.<br />

A <strong>to</strong>tal 183 samples were collected for contamination analyses, including samples<br />

<strong>of</strong> the product at different stages during manufacture (n = 136) and<br />

environmental samples (n = 47) covering most <strong>of</strong> the manufacturing surfaces during<br />

a 3-month period. Overall, the prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in raw pork before<br />

processing was 18 %. The prevalence <strong>of</strong> L. monocy<strong>to</strong>genes in pork increased<br />

<strong>to</strong> 60 % after being in brining area and brining machine. Two different L. monocy<strong>to</strong>genes<br />

PFGE types belonging <strong>to</strong> serotypes 1/2a and 1/2c were recovered from the<br />

different sites <strong>of</strong> the brining machine (feeding teeth and smooth surfaces), thus<br />

causing persistent contamination <strong>of</strong> RTE cold-smoked pork products over period<br />

<strong>of</strong> 5-years. In addition, brining machine harboured two PFGE types belonging <strong>to</strong><br />

serotypes 1/2a and 4b, which were found on different contamination sites: feeding<br />

teeth, smooth surfaces and spaces <strong>of</strong> the machine. Brining injections increased L.<br />

monocy<strong>to</strong>genes contamination in finished RTE cold-smoked pork products and<br />

processing environment. Brining area, and specifically, brining machine should<br />

be subjected for disassembling and extensive cleaning and disinfection <strong>to</strong> eliminate<br />

any persistent contamination with L. monocy<strong>to</strong>genes.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Effects <strong>of</strong> GRAS products on growth <strong>of</strong> Listeria monocy<strong>to</strong>genes during<br />

cold s<strong>to</strong>rage <strong>of</strong> salmon fillets<br />

McCarthy, S. and Johnson, D.<br />

Food and Drug Administration, USA<br />

Microbial growth can affect the quality and safety <strong>of</strong> prepared, processed, and<br />

ready-<strong>to</strong>-eat (RTE) meat, poultry, and seafood. Pathogenic bacteria, including Listeria<br />

monocy<strong>to</strong>genes (Lm), can be resistant <strong>to</strong> preservatives, sanitizers, and antibiotics.<br />

The presence <strong>of</strong> Lm on RTE products results from cross-contamination<br />

from contact surfaces in food processing plants. This study examined the ability <strong>of</strong><br />

three GRAS products <strong>to</strong> inhibit the growth <strong>of</strong> Lm on salmon fillets during cold<br />

s<strong>to</strong>rage. Twenty-five-g portions <strong>of</strong> raw salmon fillets were inoculated externally<br />

with six log 10 CFU Lm 1a1/g. Inoculated fillets were incubated at 25 °C for 20 minutes<br />

and at 4 °C for two hours <strong>to</strong> allow attachment. Filets were then treated with<br />

sterile deionized water (SDW; control) and 3 GRAS products (IONAL TM LC, MOstatin<br />

TM VS, levulinic acid; 1:5 w/v) at 25 °C for 5 min. Control and treated filets<br />

were analyzed immediately after exposure (T0) and after s<strong>to</strong>rage at 4 °C for up <strong>to</strong><br />

12 weeks. Fillets were shaken in wash buffer and cell counts in washates were determined<br />

by spread plating on chromogenic agar. Numbers <strong>of</strong> Lm increased by<br />

one log 10 during 9 weeks <strong>of</strong> cold s<strong>to</strong>rage <strong>of</strong> fillets treated with SDW. In contrast,<br />

MOstatin TMVS reduced the numbers <strong>of</strong> Lm by two logs 10 during s<strong>to</strong>rage for 9<br />

weeks. IONAL TMLC and levulinic acid prevented the growth <strong>of</strong> Lm for up <strong>to</strong> four<br />

weeks. Each <strong>of</strong> the GRAS products reduced densities <strong>of</strong> salmon bacterial flora by<br />

two <strong>to</strong> four logs 10 at two weeks compared <strong>to</strong> SDW; however, no difference was observed<br />

at 12 weeks. The use <strong>of</strong> GRAS products could prevent an increase in numbers<br />

<strong>of</strong> Lm that are associated with cross-contamination in processing plants.<br />

Heavy-metal and detergent resistance <strong>of</strong> Listeria species isolates from<br />

milk processing environments<br />

Doijad, S.*, Garg, S. and Barbuddhe, S. B.<br />

ICAR research complex for Goa, India<br />

Listeria monocy<strong>to</strong>gens is an important food-borne pathogen responsible for varied<br />

clinical forms in animal and humans. The resistance <strong>of</strong> Listeria strains <strong>to</strong> cadmium,<br />

arsenic and quaternary ammonium compounds was studied and it was correlated<br />

with resistance <strong>to</strong> quaternary ammonium compounds used as disinfectants in the<br />

food-processing industry. Limited information is available on the prevalence <strong>of</strong> resistance<br />

among isolates from the environment <strong>of</strong> food-processing plants. In this<br />

study, a <strong>to</strong>tal <strong>of</strong> 27 Listeria species isolates from the milk processing plants were<br />

included. Out <strong>of</strong> 27, 14 (51 %) found <strong>to</strong> be Listeria monocy<strong>to</strong>genes, 12 (44 %) were L.<br />

innocua and 1 (3 %) isolate was <strong>of</strong> L. ivanovii. All the isolates were subjected for determination<br />

<strong>of</strong> the resistance <strong>to</strong> cadmium, arsenic and quaternary ammonium disinfectant<br />

(benzalkonium chloride (BC)). Isolates were not inhibited at 40µg/ml <strong>of</strong><br />

cadmium chloride, while 12 (44 %) <strong>of</strong> them grew up<strong>to</strong> 200µg/ml. Isolates grew well<br />

in 40µg/ml <strong>of</strong> sodium arsenite while 3 <strong>of</strong> them could <strong>to</strong>lerate 200µg/ml. The isolates<br />

could grew at 1µg/ml <strong>of</strong> BC while only 7 could <strong>to</strong>lerate 5µg/ml. The amount <strong>of</strong><br />

BC <strong>to</strong>lerated by these isolates are well above the amount used in milk processing<br />

environment. T<strong>here</strong> was no co-relation found between resistance <strong>of</strong> isolates <strong>to</strong> cadmium,<br />

arsenate and BC, all isolates were independently resistance <strong>to</strong> metal and detergent.<br />

It was interesting <strong>to</strong> note that none <strong>of</strong> the isolates contained plasmid. This<br />

suggests that resistance against metals and detergent in Listeria spp. is not mediated<br />

by the plasmid and need further investigations. Our findings suggest that the<br />

milk processing environment constitute a reservoir for L. monocy<strong>to</strong>genes and other<br />

Listeria species. Resistance <strong>to</strong> heavy metals and quaternary ammonium disinfectant<br />

is <strong>of</strong> significance <strong>to</strong> food processing industry.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

D / P<br />

156<br />

R E F E R E N C E<br />

D / P<br />

157


R E F E R E N C E<br />

D / P<br />

158<br />

R E F E R E N C E<br />

D / P<br />

159<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Detection <strong>of</strong> Listeria monocy<strong>to</strong>genes in lettuce sold at markets<br />

and supermarkets in Por<strong>to</strong>, Portugal<br />

Noronha, L., Magalhães, R., Silva, J. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

Listeriosis is a severe infection caused by Listeria monocy<strong>to</strong>genes particularly<br />

among the elderly, very young and immunocompromized individuals and has also<br />

been associated with late-term miscarriages in pregnant women. The high incidence<br />

<strong>of</strong> L. monocy<strong>to</strong>genes in foods and the high fatality rate associated with listeriosis,<br />

has contributed <strong>to</strong> L. monocy<strong>to</strong>genes being considered a public health hazard<br />

and a continuing source <strong>of</strong> loss <strong>to</strong> food processors due <strong>to</strong> the large number <strong>of</strong> voluntary<br />

and obliga<strong>to</strong>ry recalls. Listeriosis outbreaks have been linked <strong>to</strong> the consumption<br />

<strong>of</strong> raw vegetables including lettuce. According <strong>to</strong> the European Food<br />

Standards Agency, lettuce supports L. monocy<strong>to</strong>genes growth and can be related<br />

with the transmission <strong>of</strong> listeriosis. Twenty different lettuce samples collected in<br />

markets or supermarkets located in the North <strong>of</strong> Portugal, were analyzed for the<br />

presence <strong>of</strong> L. monocy<strong>to</strong>genes by the VIDAS method. Of the tested lettuce samples,<br />

30 % were positive for the presence <strong>of</strong> L. monocy<strong>to</strong>genes. Three isolates were collected<br />

from each <strong>of</strong> the positive samples. They were serotyped by Multiplex PCR<br />

and it was possible <strong>to</strong> distinguish two serogroups: 33 % belonged <strong>to</strong> serogroup 4b –<br />

4d – 4e and 67 % <strong>to</strong> 1/2c – 3c. The present results reinforce the importance <strong>of</strong><br />

washing raw vegetables thoroughly before eating and preparing green salads and<br />

vegetable dishes shortly before eating in order <strong>to</strong> reduce the risk <strong>of</strong> listeriosis.<br />

Preliminary analysis <strong>of</strong> structure and chemical composition <strong>of</strong> extracellular<br />

polymeric substance produced by Listeria monocy<strong>to</strong>genes<br />

Nwaiwu, O.*, Lad, M., Davis, A., Foster, T. and Rees, C.<br />

Univeristy <strong>of</strong> Nottingham, UK<br />

It is generally believed that, while Listeria monocy<strong>to</strong>genes is capable <strong>of</strong> forming<br />

bi<strong>of</strong>ilms, it does not produced extensive amounts <strong>of</strong> extracellular polymer substances<br />

(EPS) that contribute <strong>to</strong> adhesion and persistence in the environment. We<br />

present <strong>here</strong> evidence that under specific growth conditions L. monocy<strong>to</strong>genes is<br />

capable EPS synthesis and this does promote surface adhesion. After EPS extraction,<br />

analysis by NMR, ATR-FTIR and SEM indicated that the material is high molecular<br />

weight and the structure <strong>of</strong> Listeria EPS is different from both polysaccharide<br />

and poly-g-glutamic acid (an EPS produced by a range <strong>of</strong> Gram-positive<br />

bacteria). Further amino acid analysis also ruled out this material being a polymer<br />

<strong>of</strong> other amino acids, while elemental energy dispersion x-ray analysis showed<br />

very low nitrogen content with carbon and oxygen the predominant elements. The<br />

as yet unidentified polymer has unusual physical properties, and scanning electron<br />

microscopy (SEM) showed rapid absorption <strong>of</strong> moisture by the EPS as relative<br />

humidity increased. Hence it is likely that this polymer will also contribute <strong>to</strong> desiccation<br />

<strong>to</strong>lerance <strong>of</strong> the bacterium.<br />

* Participation Supported by IUFoST.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Ecology and persistence <strong>of</strong> Listeria monocy<strong>to</strong>genes strains in fermented<br />

meat sausage processors from the Northern region <strong>of</strong> Portugal<br />

Ferreira, V. 1, Barbosa, J. 1, Vongkamjan, K. 2, Moreno Switt, A. 2, Hogg, T. 1, Gibbs, P. 1,<br />

Wiedmann, M. 2 and Teixeira, P. 1<br />

1. CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

2. Cornell University, USA<br />

In this study, 202 Listeria monocy<strong>to</strong>genes isolates were recovered from product<br />

samples representing 7 processors <strong>of</strong> Alheira, a traditional fermented meat sausage<br />

from the Northern region <strong>of</strong> Portugal. Samples were collected from each processor<br />

in separate dates, either at retail establishments or at processing plants. Isolates<br />

were characterized by molecular serotyping and DNA macrorestriction analysis<br />

by pulsed field gel electrophoresis (PFGE). Characterization by PFGE suggested<br />

persistence <strong>of</strong> particular strains over time in the 7 different processors, the apparent<br />

time <strong>of</strong> persistence ranged from 14 <strong>to</strong> 32 months. While a number <strong>of</strong> studies<br />

have evaluated associations <strong>of</strong> different strain characteristics, including bi<strong>of</strong>ilm<br />

formation etc., with strain persistence in processing plants, no convincing and<br />

consistent evidence for specific strain characteristics that facilitate establishment<br />

<strong>of</strong> persistence have been found so far. We thus investigated whether phage susceptibility<br />

and presence <strong>of</strong> lysogenic prophages in L. monocy<strong>to</strong>genes may be associated<br />

with strain persistence in the environment. A subset <strong>of</strong> 41 L. monocy<strong>to</strong>genes<br />

isolates representing sporadic and persistent PFGE patterns was screened for<br />

lysogeny. Twenty six prophages w<strong>here</strong> induced and their lytic spectrum against<br />

the 41 strains <strong>of</strong> L. monocy<strong>to</strong>genes was investigated. Lysogens included both sporadic<br />

and persistent PFGE types. While molecular serogroup D (4b, 4d, and 4e)<br />

isolates were more susceptible <strong>to</strong> phages as compared <strong>to</strong> serogroup A (1/2a and<br />

3a) or B (1/2b, 3b, and 7) isolates, t<strong>here</strong> was no evidence for differences in phage<br />

susceptibility between persistent and sporadic strains. While our findings support<br />

that L. monocy<strong>to</strong>genes serotypes differ in phage resistance, strain persistence does<br />

not seem <strong>to</strong> be associated with enhanced phage resistance.<br />

Bi<strong>of</strong>ilm formation and survival <strong>of</strong> L. monocy<strong>to</strong>genes and slaughter house<br />

bacteria on surfaces at relevant environmental conditions<br />

Langsrud, S., Møretrø, T. and Heir, E.<br />

Norwegian Institute <strong>of</strong> Food, Fisheries and Aquaculture Research, Norway<br />

Adhesion, bi<strong>of</strong>ilm formation and survival <strong>of</strong> microorganisms on food contact surfaces<br />

are <strong>of</strong> concern in the food processing industry. Surface associated bacteria<br />

can lead <strong>to</strong> cross contamination <strong>of</strong> food and have serious consequences for human<br />

health and food quality. Persistence <strong>of</strong> Listeria monocy<strong>to</strong>genes in the food processing<br />

environment is a well known food safety concern. Also, other environmental<br />

bacteria may cause problems when persisting in the food industry. A better<br />

knowledge on the ability <strong>of</strong> dominant bacteria <strong>to</strong> produce bi<strong>of</strong>ilm and survive on<br />

food contact surfaces under relevant conditions is t<strong>here</strong>fore important. We have<br />

studied bi<strong>of</strong>ilm formation and survival <strong>of</strong> L. monocy<strong>to</strong>genes and environmental<br />

bacteria isolated from surfaces in a meat slaughter house. Bacteria were isolated<br />

after cleaning and disinfection and prior <strong>to</strong> slaughter on a regular working day<br />

and identified by 16S rDNA sequencing. Selected isolates from dominating genera<br />

and L. monocy<strong>to</strong>genes were tested for their ability <strong>to</strong> form bi<strong>of</strong>ilms under various<br />

temperature conditions in a microtiter plate assay. We also studied survival on<br />

stainless steel under controlled temperature and humidity conditions. The susceptibility<br />

<strong>of</strong> surface ad<strong>here</strong>d bacterial cells <strong>to</strong> four commercial disinfectants commonly<br />

used in the meat industry was investigated. Predominant bacteria in the<br />

slaughtering line belonged <strong>to</strong> the genera Pseudomonas, Serratia, Acine<strong>to</strong>bacter,<br />

Citrobacter, Aerococcus, Kocuria and Staphylcococcus. The data indicated low<br />

bi<strong>of</strong>ilm forming abilities <strong>of</strong> L. moncy<strong>to</strong>genes at relevant food industry temperatures,<br />

while other environmental bacteria showed variable bi<strong>of</strong>ilm formation at<br />

12 and 20 °C. Gram-positive bacteria showed excellent survival on stainless steel<br />

during incubation at 12 °C, 70 % relative humidity. The bactericidal effects <strong>of</strong> disinfectants<br />

on bacteria ad<strong>here</strong>d <strong>to</strong> surfaces were low (< 3 log kill) for all tested disinfectants.<br />

In conclusion, survival <strong>of</strong> bacteria after cleaning and disinfection in<br />

the meat industry can be explained by resistance <strong>to</strong> disinfectants when dried on<br />

surfaces and either ability <strong>to</strong> survive at dry conditions or bi<strong>of</strong>ilm formation in the<br />

presence <strong>of</strong> water.<br />

R E F E R E N C E<br />

D / P<br />

160<br />

R E F E R E N C E<br />

D / P<br />

161


R E F E R E N C E<br />

D / P<br />

162<br />

R E F E R E N C E<br />

D / P<br />

163<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Control <strong>of</strong> L. monocy<strong>to</strong>genes by lysozyme combined with olive leaf<br />

extract in edible pullulan film coated on chicken breast fillets<br />

Handan Baysal, A.*<br />

Izmir Institute <strong>of</strong> Technology, Department <strong>of</strong> Food Engineering, Urla, Izmir, Turkey<br />

In recent years, naturally occurring antimicrobial and antioxidant compounds<br />

have been preferably employed in meats because <strong>of</strong> their potential health benefits<br />

and safety compared with synthetic preservatives. The objective <strong>of</strong> this study was<br />

<strong>to</strong> evaluate the antimicrobial effect <strong>of</strong> olive leaf extract (OLE) and lysozyme added<br />

<strong>to</strong> pullulan film coatings against Listeria monocy<strong>to</strong>genes in raw chicken breast meat<br />

s<strong>to</strong>red aerobically under refrigeration (4 °C). The effectiveness <strong>of</strong> these compounds<br />

in a meat model system was evaluated by surface inoculation (approximately<br />

10 6 CFU/g) <strong>of</strong> L. monocy<strong>to</strong>genes on<strong>to</strong> chicken breast fillets. Fresh chicken<br />

breast fillets used in this research was purchased from a local supermarket. The inoculated<br />

raw chicken breast fillets were treated before s<strong>to</strong>rage by dipping in<strong>to</strong> 5 %<br />

(w/v) pullulan film-forming solutions with and without the addition <strong>of</strong> antimicrobial<br />

agents: (a) deionized water, (b) 5 % (w/v) pullulan + 20 % (w/v) water extract<br />

<strong>of</strong> OLE, (c) 5 % (w/v) pullulan + lysozyme (d) 5 % (w/v) pullulan + 20 % (w/v)<br />

water extract <strong>of</strong> OLE + lysozyme) for 10 min at 20 °C. After removal <strong>of</strong> excess moisture,<br />

the samples were s<strong>to</strong>red at 4 °C. The inhibi<strong>to</strong>ry effects <strong>of</strong> OLE and lysozyme<br />

containing edible coatings were evaluated for 14 d. In the meat system, the L.<br />

monocy<strong>to</strong>genes population was decreased effectively after 14 d at 4 °C. This research<br />

has demonstrated that the use <strong>of</strong> an edible film coating containing natural<br />

extracts or the application <strong>of</strong> a lysozyme solution is a promising means <strong>of</strong> controlling<br />

the growth and recontamination <strong>of</strong> L. monocy<strong>to</strong>genes on raw chicken<br />

breast meat s<strong>to</strong>red under refrigeration.<br />

* Participation Supported by IUFoST.<br />

Evaluation <strong>of</strong> antilisterial activity by lactic acid bacteria<br />

Borges, S., Barbosa, J., Albano, H., Silva, J. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

Listeriosis is an infection with a high morbidity and mortality, caused by Listeria<br />

monocy<strong>to</strong>genes. It occurs primarily in elderly patients, immunocompromised individuals,<br />

pregnant women and their neonates. Some lactic acid bacteria (LAB)<br />

are able <strong>to</strong> produce antimicrobial compounds with activity against several<br />

pathogens, such as organic acids (lactic and acetic acids), hydrogen peroxide, antimicrobial<br />

enzymes and bacteriocins. The objective <strong>of</strong> this work was <strong>to</strong> evaluate<br />

the antimicrobial activity <strong>of</strong> selected LAB against L. monocy<strong>to</strong>genes. Thirty-five<br />

isolates <strong>of</strong> LAB, available at the culture collection <strong>of</strong> Escola Superior de Biotecnologia,<br />

demonstrated inhibi<strong>to</strong>ry activity against 29 clinical isolates <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

To characterize this antilisterial activity the antagonistic spectrum <strong>of</strong> each<br />

LAB culture, neutralized cell-free supernant and neutralized cell-free supernant<br />

treated with catalase or trypsin were investigated. All isolates <strong>of</strong> LAB showed antibacterial<br />

effect probably due <strong>to</strong> the production <strong>of</strong> bacteriocins. The bacteriocin<br />

activity (AU/mL) was determinated for three serotypes <strong>of</strong> L. monocy<strong>to</strong>genes (1/2a,<br />

1/2b, 4b) and the activity varied between 800 and 6400 AU/mL. According <strong>to</strong> this<br />

study, the use <strong>of</strong> bacteriogenic LAB can be important <strong>to</strong> control L. monocy<strong>to</strong>genes<br />

<strong>of</strong> clinical origin.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Molecular methods <strong>to</strong> assess Listeria monocy<strong>to</strong>genes route<br />

<strong>of</strong> contamination in a dairy processing plant<br />

Cocolin, L., Alessandria, V., Dolci, P. and Rantsiou, K.<br />

Univeristy <strong>of</strong> Turin, Italy<br />

In this study we investigated the occurrence <strong>of</strong> Listeria monocy<strong>to</strong>genes in a dairy<br />

processing plant during two sampling campaigns in 2007 and 2008. Samples represented<br />

by semi-finished and finished cheeses, swabs from the equipment and<br />

brines from the salting step, were subjected <strong>to</strong> analysis by using traditional and<br />

molecular methods, represented mainly by quantitative PCR. Comparing the results<br />

obtained by the application <strong>of</strong> the two approaches used, it became evident<br />

how traditional microbiological analysis underestimated the presence <strong>of</strong> L. monocy<strong>to</strong>genes<br />

in the dairy plant. Especially samples <strong>of</strong> the brines and the equipment<br />

swabs were positive only with qPCR. For some equipment swabs it was possible <strong>to</strong><br />

detect a load <strong>of</strong> 10 4-10 5 cfu/cm 2, while the ISO method employed gave negative<br />

results both before and after the enrichment step. The evidences collected during<br />

the first sampling year, highlighting a heavy contamination <strong>of</strong> the brines and <strong>of</strong><br />

the equipment, lead <strong>to</strong> the implementation <strong>of</strong> specific actions that decreased the<br />

contamination in these samples during the 2008 campaign. However, no reduction<br />

in the number <strong>of</strong> L. monocy<strong>to</strong>genes positive final products was observed, suggesting<br />

that a more strict control is necessary <strong>to</strong> avoid the presence <strong>of</strong> the<br />

pathogen. All the isolates <strong>of</strong> L. monocy<strong>to</strong>genes were able <strong>to</strong> form bi<strong>of</strong>ilm, and, interestingly,<br />

considering the results obtained from their molecular characterization<br />

it became evident how strains present in the brines, were genetically connected<br />

with isolates from the equipment and from the final product, suggesting a<br />

clear route <strong>of</strong> contamination <strong>of</strong> the pathogen in the dairy plant. This study underlines<br />

the necessity <strong>to</strong> use appropriate analytical <strong>to</strong>ols, such as molecular methods,<br />

<strong>to</strong> fully understand the spread and persistence <strong>of</strong> L. monocy<strong>to</strong>genes in food producing<br />

companies.<br />

Persistence <strong>of</strong> L. monocy<strong>to</strong>genes in artisanal cheese producing plants<br />

Almeida, G., San<strong>to</strong>s, I., Magalhães, R., Barbosa, J., Hogg, T. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

The consumption <strong>of</strong> raw milk or raw milk products has caused several listeriosis<br />

outbreaks resulting in several hundred cases. This highlights the risk <strong>of</strong> these products<br />

and led public health <strong>of</strong>ficials <strong>to</strong> recommend that raw milk and dairy products<br />

prepared from raw milk should not be consumed by susceptible populations, particularly<br />

pregnant women. The presence <strong>of</strong> L. monocy<strong>to</strong>genes in cheeses has been<br />

widely reported in the literature over the years. The prevention <strong>of</strong> Listeria contamination<br />

is difficult <strong>to</strong> achieve in raw milk cheeses. Also, contaminated raw milk<br />

can contaminate the plant environment and some strains can colonize and persist<br />

for long periods. The present work aimed <strong>to</strong> characterize L. monocy<strong>to</strong>genes isolates<br />

recovered between 2004 and 2007 from an artisanal ewe’s raw milk cheese producing<br />

plant. PFGE using restriction enzymes AscI and ApaI was performed in<br />

forty isolates: 18 from cheese, one from raw milk, 20 from environmental sites, and<br />

one from whey. Six combined pulsotypes were obtained, one <strong>of</strong> them aggregating 20<br />

isolates. This is an evidence <strong>of</strong> the presence <strong>of</strong> a resident clone with a persistence<br />

time estimated in at least, 14 months. The pr<strong>of</strong>ile obtained from cheese isolates<br />

was also encountered in isolates from ewe’s raw milk and from raw milk reception’s<br />

floor suggesting that raw milk could be the source <strong>of</strong> contamination. Control L.<br />

monocy<strong>to</strong>genes in the environment plays an important role in reducing its presence<br />

in foods, which is necessary <strong>to</strong> reduce the incidence <strong>of</strong> listeriosis.<br />

R E F E R E N C E<br />

D / P<br />

164<br />

R E F E R E N C E<br />

D / P<br />

165


R E F E R E N C E<br />

D / P<br />

166<br />

R E F E R E N C E<br />

D / P<br />

167<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Occurrence <strong>of</strong> Listeria monocy<strong>to</strong>genes in food products collected in<br />

Portugal from retail establishments and food plants<br />

Mena, C., Carneiro, L., San<strong>to</strong>s, I., Magalhães, R., Almeida, G. and Teixeira, P.<br />

CBQF / Universidade Católica Portuguesa – Escola Superior de Biotecnologia, Por<strong>to</strong>, Portugal<br />

Listeriosis is an infection caused by the bacterium Listeria monocy<strong>to</strong>genes. The<br />

ingestion <strong>of</strong> contaminated food with this microorganism may cause serious health<br />

problems for consumers. During two years (2007 and 2008) the presence <strong>of</strong> L<br />

monocy<strong>to</strong>genes was evaluated in a <strong>to</strong>tal <strong>of</strong> 1476 food samples, collected from retail<br />

and food plants. The detection <strong>of</strong> the microorganism was performed using the au<strong>to</strong>mated<br />

VIDAS system and the positive results were confirmed following the ISO<br />

11290 standard. L. monocy<strong>to</strong>genes was detected in 134 (9.1 %) <strong>of</strong> the analyzed samples.<br />

Most <strong>of</strong> positive samples were from foods normally submitted <strong>to</strong> heat treatment<br />

before consumption such as pré-cooked foods (30.0 %, 12/40; pizza, pasta,<br />

rissois, etc) and fermented meat products (20.6 %, 33/160; farinheira, alheira,<br />

morcela, bacon), from raw products (16.0 %, 58/362; raw meat, vegetables and fish),<br />

and from “ready-<strong>to</strong>-eat” foods: fermented meat products (8.8 %, 10/114), vegetables<br />

salads (1.1 %, 2/174), ready cooked meals (4.5 %, 6/134), cheeses (4.6 %, 12/262)<br />

and fresh cheese (1.3 %, 1/80). The occurrence <strong>of</strong> the microorganism in ready-<strong>to</strong>eat<br />

foods is <strong>of</strong> more concern. L. monocy<strong>to</strong>genes grows at refrigeration temperatures<br />

and could achieve levels <strong>of</strong> contamination that can cause disease. Also, foods<br />

that will have a heat treatment at consumer’s home could represent a hazard if<br />

cross contamination <strong>of</strong> food items that will be consumed without any further step<br />

<strong>of</strong> destruction occurs.<br />

Listeria monocy<strong>to</strong>genes bi<strong>of</strong>ilms grown at 12 °C showed reduced<br />

susceptibility <strong>to</strong> sanitizers<br />

Lourenço, A., Machado, H.* and Bri<strong>to</strong>, L.<br />

CBAA – Departamen<strong>to</strong> de Botânica e Engenharia Biológica, Institu<strong>to</strong> Superior de Agronomia, Technical<br />

University <strong>of</strong> Lisbon, Portugal<br />

Listeria monocy<strong>to</strong>genes may form bi<strong>of</strong>ilms on food contact surfaces <strong>of</strong> difficult sanitization<br />

which may lead <strong>to</strong> recurrent contamination <strong>of</strong> food products. The eradication<br />

<strong>of</strong> bi<strong>of</strong>ilms can only be achieved by using adequate hygienization routines<br />

which will ultimately ensure food safety. The bi<strong>of</strong>ilm forming ability <strong>of</strong> four L.<br />

monocy<strong>to</strong>genes strains from different origins, cheese, dairy environment and from<br />

human cases <strong>of</strong> listeriosis was evaluated, either in pure culture or in co-culture<br />

with Pseudomonas aeruginosa, at 37 °C and 12 °C using the Calgary Bi<strong>of</strong>ilm Device®<br />

(CBD). The minimum bi<strong>of</strong>ilm eradication concentration (MBEC) was determined<br />

for four commercial dairy sanitizers (one alkyl amine acetate based, T99;<br />

two chlorine based, T66 and DD and one phosphoric acid based, BP). Co-culture<br />

bi<strong>of</strong>ilms had an average <strong>to</strong>tal population <strong>of</strong> 7 <strong>to</strong> 8 Log 10 CFU/peg. P.aeruginosa<br />

was the dominant species, either at 37 °C or at 12 °C, representing 99 % <strong>of</strong> the <strong>to</strong>tal<br />

CFU/peg. L. monocy<strong>to</strong>genes bi<strong>of</strong>ilms grown, either at 37 °C or 12 °C, although with<br />

different incubation times (24 hours and 7 days, respectively) reached a similar<br />

cell density (6 Log 10 CFU/peg). Nevertheless, the bi<strong>of</strong>ilms produced at 12 °C were<br />

generally less susceptible <strong>to</strong> the sanitizers than when produced at 37 °C. One <strong>of</strong><br />

the strains (3880) retrieved MBEC values <strong>of</strong> 30720 µg/ml and 16000 µg/ml for<br />

T99 and BP, respectively. These values were above the maximum in-use recommended<br />

concentrations (T99 – 29700 µg/ml and BP – 11500 µg/ml) for these<br />

agents. The growth in co-culture also proved <strong>to</strong> be relevant regarding disinfectant<br />

susceptibility, as the co-cultures were generally less susceptible than L. monocy<strong>to</strong>genes<br />

pure cultures. The MBEC values obtained for the chlorine based agents<br />

were never over recommended in-use concentrations which may indicate a more<br />

efficient ability <strong>to</strong> eradicate bi<strong>of</strong>ilms, if in-use conditions are met.<br />

* Participation Supported by IUFoST.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Modelling growth <strong>of</strong> Listeria monocy<strong>to</strong>genes in cheese as function<br />

<strong>of</strong> environmental variables<br />

Sand Rosshaug, P. and Hallberg Larsen, M.<br />

Faculty <strong>of</strong> Life Science, Department <strong>of</strong> Veterinary Pathobiology, Denmark<br />

This project is part <strong>of</strong> a larger project with the purpose <strong>to</strong> develop a predictive<br />

model <strong>of</strong> growth <strong>of</strong> the pathogen Listeria monocy<strong>to</strong>genes in a dairy chain producing<br />

blue/white mould cheese. This project focuses on the growth kinetics <strong>of</strong> L. monocy<strong>to</strong>genes<br />

as function <strong>of</strong> environmental variables. Temperature, water activity, pH,<br />

lactic acid, NaCl, and O 2 are environmental fac<strong>to</strong>rs that impact the growth kinetics<br />

<strong>of</strong> L. monocy<strong>to</strong>genes. The growth kinetics was investigated for 3 strains <strong>of</strong> L.<br />

monocy<strong>to</strong>genes that has been found in the dairy industry: ATCC 19111, ATCC 19113,<br />

ATCC 19115. Different growth media were tested: broth, milk, and blue/white<br />

mould cheese. The growth kinetics <strong>of</strong> L. monocy<strong>to</strong>genes as function <strong>of</strong> these variables<br />

was formulated as an ordinary differential equation using Cardinal Parameters<br />

<strong>to</strong> describe the influence <strong>of</strong> the environmental fac<strong>to</strong>rs. The growth kinetics<br />

was applied in a predictive mathematical, deterministic model <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

Finally the growth kinetics <strong>of</strong> L. monocy<strong>to</strong>genes was investigated using a<br />

s<strong>to</strong>chastic predictive model taking in<strong>to</strong> account the s<strong>to</strong>chastic nature <strong>of</strong> some <strong>of</strong><br />

the inputs including the lag.<br />

Characterization <strong>of</strong> anti-Listerial bacteriocin produced by Lac<strong>to</strong>bacillus<br />

plantarum ST8SH, a strain isolated from Bulgarian salami<br />

Todorov, S. D. 1* and Lemos Vaz-Velho, M. 2<br />

1. Universidade de São Paulo, Faculdade de Ciências Farmacêuticas, Brazil<br />

2. Escola Superior de Tecnologia e Gestão, Institu<strong>to</strong> Politécnico de Viana do Castelo, Portugal<br />

Strain ST8SH, isolated from Bulgarian salami, was identified as Lac<strong>to</strong>bacillus plantarum<br />

based on biochemical tests, sugar fermentation reactions (API50CHL), PCR<br />

with species-specific primers and 16S rDNA sequencing. Strain ST8SH produces a<br />

3.0kDa class IIa bacteriocin, active against Listeria monocy<strong>to</strong>genes, Listeria innocua,<br />

Strep<strong>to</strong>coccus caprinus, Strep<strong>to</strong>coccus spp., Lac<strong>to</strong>bacillus casei, Lac<strong>to</strong>bacillus<br />

curvatus, Lac<strong>to</strong>bacillus salivarius, Lac<strong>to</strong>bacillus pen<strong>to</strong>sus, Enterococcus mundtii,<br />

Enterococcus faecalis and Lac<strong>to</strong>coccus lactis subsp. lactis. No change in activity<br />

was recorded after 2 h at pH values between 2.0 and 12.0, and after treatment at<br />

100 °C for 120 min or 121 °C for 20 min. The mode <strong>of</strong> activity against L. innocua, L.<br />

monocy<strong>to</strong>genes is bactericidal, resulting in cell lyses and enzyme- and DNA-leakage<br />

and was visualised by a<strong>to</strong>mic force microscopy. The highest level <strong>of</strong> activity (25600<br />

AU/ml) was recorded when cells were grown at 37 °C or 30 °C in MRS broth (pH<br />

6.5). Peptide ST8SH adsorbs at low levels (400 AU/ml) <strong>to</strong> producer cells. High cell<br />

numbers <strong>of</strong> L. plantarum ST8SH and L. innocua LMG13568 were recorded at beginning<br />

when co-cultured. However, the cell numbers <strong>of</strong> L. innocua LMG13568 decreased<br />

from 1.6x10 4 CFU/ml <strong>to</strong> 2.5x10 2 CFU/ml in 12 h and <strong>to</strong> undetectable levels<br />

after 24 h. Plantaricin ST8SH production was stimulated by presence <strong>of</strong> L. innocua<br />

LMG13568 (102 400 AU/ml). Similar results were obtained with addition <strong>of</strong> 10 %<br />

au<strong>to</strong>claved overnight culture <strong>of</strong> L. innocua LMG13568 <strong>to</strong> MRS growth media on<br />

production <strong>of</strong> plantaricin ST8SH. Based on the genetic approach strain ST8SH<br />

harbours associated genetic determinants for production <strong>of</strong> a variation <strong>of</strong> the well<br />

known plantaricin 423. Future purification <strong>of</strong> the produced bacteriocin need <strong>to</strong><br />

be performed <strong>to</strong> determine if Lac<strong>to</strong>bacillus plantarum ST8SH produces this bacteriocin<br />

(plantaricin 423 – like) or harbour more then one bacteriocin operons.<br />

* Participation Supported by IUFoST.<br />

R E F E R E N C E<br />

D / P<br />

168<br />

R E F E R E N C E<br />

D / P<br />

169


R E F E R E N C E<br />

D / P<br />

170<br />

R E F E R E N C E<br />

D / P<br />

171<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

EFSA’s proposal for an EU-wide retail survey on Listeria monocy<strong>to</strong>genes<br />

in selected categories <strong>of</strong> ready-<strong>to</strong>-eat food products<br />

Boelaert, F., Felício, T. and Makela, P.<br />

EFSA, Italy<br />

The European Food Safety Authority and its Task Force on Zoonoses Data Collection<br />

were requested by the European Commission <strong>to</strong> provide a proposal for technical<br />

specifications on an EU-wide retail survey on Listeria monocy<strong>to</strong>genes in selected<br />

categories <strong>of</strong> ready-<strong>to</strong>-eat (RTE) food products that should take place during<br />

the whole year <strong>of</strong> 2010. The proposed technical specifications focus on sampling<br />

those categories <strong>of</strong> RTE food in which the highest L. monocy<strong>to</strong>genes contamination<br />

have been observed in the European Union (EU): s<strong>of</strong>t and semi-s<strong>of</strong>t cheeses,<br />

smoked and gravad fish, and heat-treated meat products that are handled after<br />

heat treatment. Sampling <strong>of</strong> these RTE food categories would be targeted at retail<br />

outlets serving the final consumer, with catering and wholesale establishments<br />

excluded. Food products are suggested <strong>to</strong> be tested at the end <strong>of</strong> the shelf-life and<br />

additionally in the case <strong>of</strong> smoked and gravad fish, immediately after sampling. At<br />

the Community-level a <strong>to</strong>tal <strong>of</strong> about 3,000 samples should be taken for each RTE<br />

food category, per analyses stage. A specific number <strong>of</strong> samples is proposed <strong>to</strong> be<br />

proportionally allocated <strong>to</strong> the Member States according <strong>to</strong> the size <strong>of</strong> their human<br />

populations. Standardised analytical L. monocy<strong>to</strong>genes detection and enumeration<br />

methods are proposed <strong>to</strong> be employed in the analyses <strong>of</strong> samples. In addition,<br />

water activity and pH values are <strong>to</strong> be measured in the smoked and gravad fish.<br />

This proposal for a Community-specific survey would only allow estimation <strong>of</strong> the<br />

L. monocy<strong>to</strong>genes prevalence at the Community level. A modelling and simulation<br />

approach should be applied in the analyses <strong>of</strong> the results so that the effectiveness<br />

<strong>of</strong> the implementation <strong>of</strong> Community L. monocy<strong>to</strong>genes criteria may be assessed. A<br />

similar model-based approach will also be used <strong>to</strong> estimate the growth potential <strong>of</strong><br />

L. monocy<strong>to</strong>genes in smoked and gravad fish.<br />

Ripening conditions: an asset <strong>to</strong> control L. monocy<strong>to</strong>gnes in cheeses<br />

Callon, C., Picque, D., Corrieu, G. and Montel, M.-C.<br />

INRA, France<br />

The EC regulations for dairy products require the absence <strong>of</strong> L. monocy<strong>to</strong>genes<br />

output production with a possible derogation at < 100cfu/g if it is demonstrated<br />

that t<strong>here</strong> is no evolution during s<strong>to</strong>rage until consumption. To identify the microbial<br />

populations <strong>of</strong> cheeses and environmental fac<strong>to</strong>rs that may be barrier <strong>to</strong> L.<br />

monocy<strong>to</strong>genes in the European project Truefood an ecological approach was preferred<br />

over a process <strong>of</strong> screening <strong>of</strong> strains. This approach based on the following<br />

steps: 1) selection <strong>of</strong> milk with anti-Listeria properties, 2) identification <strong>of</strong> microbial<br />

communities and simplifications <strong>of</strong> their composition, 3) cheese-making experiments<br />

in different conditions <strong>of</strong> ripening. This strategy was successfully applied<br />

<strong>to</strong> the inhibition <strong>of</strong> L. monocy<strong>to</strong>genes in cheese core. Indeed, a microbial<br />

community <strong>of</strong> raw milk, composed <strong>of</strong> 7 strains <strong>of</strong> lactic acid bacteria (Lac<strong>to</strong>bacillus<br />

and Leuconos<strong>to</strong>c), in synergy with bacteria called ripening Gram positive (Staphylococcus,<br />

Arthrobacter, Brachybacterium, Microbacterium, Corynebacterium, Brevibacterium)<br />

inhibited L. monocy<strong>to</strong>genes at the same extent than the complex community.<br />

The decrease <strong>of</strong> relative humidity from 98 <strong>to</strong> 93 % can also act as a barrier<br />

against L. monocy<strong>to</strong>genes, especially at the surface <strong>of</strong> cheeses with or without microbial<br />

consortium. It led also <strong>to</strong> an increase <strong>of</strong> dry matter at the surface <strong>of</strong> cheeses<br />

and a decrease <strong>of</strong> pH which may contribute <strong>to</strong> the inhibition <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

Interestingly, at the beginning <strong>of</strong> ripening (8 days), a temperature <strong>of</strong> 13 °C, by<br />

favouring the growth <strong>of</strong> Lac<strong>to</strong>bacillus and Leuconos<strong>to</strong>c composing the microbial<br />

consortium, was more effective than 9 °C for inhibiting Listeria. This can be explained<br />

by the highest galac<strong>to</strong>se metabolism and acid production. On the contrary<br />

in the control with only S. thermophilus, the ripening temperature <strong>of</strong> 13 °C<br />

favoured the growth <strong>of</strong> L. monocy<strong>to</strong>genes. The ripening conditions reasoned according<br />

<strong>to</strong> the composition <strong>of</strong> microbial communities could be a promising strategy<br />

in the control <strong>of</strong> L. monocy<strong>to</strong>genes.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Deterministic and s<strong>to</strong>chastic behavior <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

suspended cells or detached from stainless steel surfaces during<br />

cheese manufacturing<br />

Belessi, C.-E. A. 1, Gounadaki, A. S. 1, Arapakh, S. 1, Schvartzman, S. 2, Jordan, K. 2<br />

and Skandamis, P. N. 1<br />

1. Agricultural University <strong>of</strong> Athens, Greece<br />

2. Teagasc, Dairy Products Research Centre, Moorepark, Fermoy, Co. Cork, Ireland<br />

Growth probability and kinetic models for Listeria monocy<strong>to</strong>genes in response <strong>to</strong><br />

multiple hurdles occurring during cheese manufacturing are mainly focused on<br />

suspended L. monocy<strong>to</strong>genes cells. This study aimed <strong>to</strong> compared: (i) the<br />

growth/no growth interface <strong>of</strong> L. monocy<strong>to</strong>genes cells attached on stainless steel<br />

(SS) surfaces, or in suspension, within adjusted media and (ii) the behavior <strong>of</strong><br />

plank<strong>to</strong>nic and detached Listeria cells during manufacturing and ripening <strong>of</strong> two<br />

popular Greek cheeses: Feta and Graviera. A multi-strains composite <strong>of</strong> L. monocy<strong>to</strong>genes<br />

isolates from cheese, fac<strong>to</strong>ry and farm in Greece and Ireland, were grown<br />

in TSBYE, MRD, Milk, Feta and Graviera cheese in the presence <strong>of</strong> SS coupons<br />

(2x5cm 2) for 3d at 20 °C, <strong>to</strong> obtain the following inocula: plank<strong>to</strong>nic cells (P), and<br />

cells detached from the SS coupons (D). Detachment <strong>to</strong>ok place by the bead vortexing<br />

method. For growth/no growth evaluation P and D cells were inoculated in<br />

TSBYE, adjusted <strong>to</strong> 5 pH (6.8-4.8) by lactic acid and at 4 a w (0.945-0.995) by NaCl.<br />

For evaluation <strong>of</strong> L. monocy<strong>to</strong>genes kinetics in cheese, P and D cells were inoculated<br />

at three simulated stages <strong>of</strong> Feta and Graviera manufacture: in pasteurized<br />

milk, after cutting the curd and after the first ripening. The growth <strong>of</strong> D cells<br />

slightly delayed compared <strong>to</strong> P cells while it was more affected by a w than pH. On<br />

cheese, L. monocy<strong>to</strong>genes survived throughout the ripening at low levels. The differences<br />

in probability <strong>of</strong> growth <strong>of</strong> single cells for both inocula (P and D) were<br />

assessed by s<strong>to</strong>chastic approaches. Furthermore, PFGE analysis resulted that 91 %<br />

<strong>of</strong> the cells <strong>of</strong> any tested condition belonged <strong>to</strong> the cheese fac<strong>to</strong>ry isolate. The results<br />

may address safety implications relevant <strong>to</strong> the potential <strong>of</strong> attached cells <strong>to</strong><br />

proliferate, w<strong>here</strong>as data may contribute <strong>to</strong> filling data gaps on risk assessment<br />

<strong>of</strong> L. monocy<strong>to</strong>genes isolates from the dairy industry.<br />

Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in game meat<br />

Atanassova, V.<br />

Institute <strong>of</strong> Food Quality and Food Safety, Germany<br />

Listeria spp. are widespread in the environment. Foods <strong>of</strong> animal origin like raw<br />

meat, milk, s<strong>of</strong>t cheese and smoked fish are <strong>of</strong>ten contaminated by Listeria. During<br />

slaughtering and food processing, these bacteria can distribute and establish as<br />

specific contaminants in difficult <strong>to</strong> clean areas <strong>of</strong> the plant and the processing<br />

environment. Freshly shot game meat can be readily contaminated during the cutting<br />

and trimming steps. Listeria endures most non heat processing steps and will<br />

be present in packaged retail meats. Chill s<strong>to</strong>rage for long periods <strong>of</strong> time can result<br />

in a considerable increase in numbers. In this study 797 (roe deer, red deer<br />

and wild boar) samples <strong>of</strong> wild game from freshly shot whole carcasses and 481<br />

samples from packaged game meat s<strong>to</strong>red under chill conditions (+ 4°C) were analyzed<br />

for the presence <strong>of</strong> Listeria spp. and Listeria monocy<strong>to</strong>genes. Standard methods<br />

according <strong>to</strong> ISO 11290 were applied. Enrichment <strong>of</strong> meat samples was performed<br />

first in half Fraser, followed by full Fraser broth. Enrichment cultures were<br />

streaked <strong>to</strong> OCLA plates. In <strong>to</strong>tal 7.3 % (n=58) <strong>of</strong> the meat from whole carcasses<br />

was positive for Listeria spp., with a <strong>to</strong>tal <strong>of</strong> 39 (4.9 %) samples being contaminated<br />

by Listeria monocy<strong>to</strong>genes. T<strong>here</strong> was no difference in the prevalence in<br />

meat from different game species. Chill s<strong>to</strong>red packaged game meat showed higher<br />

prevalence rates <strong>of</strong> 20.6 % (n=99) Listeria monocy<strong>to</strong>genes positive samples. The<br />

results show that game meat can be contaminated by Listeria monocy<strong>to</strong>genes.<br />

Prevalence was lower at the beginning <strong>of</strong> the meat processing in carcasses still unskinned,<br />

while during elongated s<strong>to</strong>rage the rate can be higher in chill s<strong>to</strong>red meat<br />

cuts. To reduce the risk <strong>of</strong> infection for the consumer, proper cooking is advised<br />

during meal preparation.<br />

R E F E R E N C E<br />

D / P<br />

172<br />

R E F E R E N C E<br />

D / P<br />

173


R E F E R E N C E<br />

D / P<br />

174<br />

R E F E R E N C E<br />

D / P<br />

175<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Relationship between pathogenic pr<strong>of</strong>ile and in vitro bi<strong>of</strong>ilm formation<br />

capacity <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated from meat, fish and<br />

processing plants<br />

Meloni, D., Mazza, R., Marceddu, M., Piras, F., Mureddu, A. and Mazzette, R.<br />

Department <strong>of</strong> Animal Biology, Faculty <strong>of</strong> Veterinary Medicine, University <strong>of</strong> Sassari<br />

In the present survey, the relationships between pathogenic pr<strong>of</strong>ile and in vitro<br />

bi<strong>of</strong>ilm formation <strong>of</strong> 106 Listeria monocy<strong>to</strong>genes strains having no epidemiological<br />

correlation and isolated from different environmental and food sources, were analyzed.<br />

The isolates were grouped in<strong>to</strong> different categories based on the source <strong>of</strong><br />

isolation: swine and poultry carcasses (14 and 13 % respectively), ground meat<br />

(7 %), fermented sausages (10 %), raw and smoked salmons (9 and 4 % respectively),<br />

swine slaughterhouse environments (3%), fermented sausage and smoked salmon<br />

processing plants environments (24 and 16 % respectively). The quantitative assessment<br />

<strong>of</strong> the in vitro bi<strong>of</strong>ilm formation was carried out by using a microtiter<br />

plate assay with spectropho<strong>to</strong>metric reading (OD 620 ). The cut-<strong>of</strong>f value (ODc) was<br />

equal <strong>to</strong> three time the standard deviation <strong>of</strong> the negative controls plus the average<br />

OD reading for the same controls. The strains were divided up in<strong>to</strong> four categories,<br />

based on their ability <strong>to</strong> form bi<strong>of</strong>ilms: no bi<strong>of</strong>ilm producers (OD≤ODc), weak producers<br />

(ODc


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Minimum Bi<strong>of</strong>ilm Eradication Concentration (MBEC)<br />

<strong>of</strong> different antimicrobials on Listeria monocy<strong>to</strong>genes<br />

and Salmonella enterica bi<strong>of</strong>ilms<br />

Rodrigues, D. 1, Teixeira, P. 1, Oliveira, R.1, Ceri, H. 2 and Azeredo, J. 1<br />

1. Institute for Biotechnology and Bioengineering, Centre <strong>of</strong> Biological Engineering, Universidade do<br />

Minho, Campus de Gualtar, Braga, Portugal<br />

2. Bi<strong>of</strong>ilm Research Group, Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Calgary, Canada<br />

Inadequate disinfection <strong>of</strong> food processing environments contributes <strong>to</strong> foodborne<br />

disease outbreaks, particularly those concerning L. monocy<strong>to</strong>genes and Salmonella<br />

enterica. The ability <strong>of</strong> these bacteria <strong>to</strong> ad<strong>here</strong> and form bi<strong>of</strong>ilms on several surfaces<br />

makes sanitation more difficult and challenging, reason why susceptibility<br />

tests must cover not only plank<strong>to</strong>nic cells but also ad<strong>here</strong>d cells and bi<strong>of</strong>ilms.<br />

Through MBEC assessment using Calgary Bi<strong>of</strong>ilm Device (CBD), this work aimed<br />

at comparing the performance <strong>of</strong> four antimicrobials on L. monocy<strong>to</strong>genes and Salmonella<br />

enterica bi<strong>of</strong>ilms and the disinfection efficiency between strains and<br />

species. Three L. monocy<strong>to</strong>genes strains (994, 1562 and CECT 4031 T) and five S.<br />

enterica strains (355, CC, NCTC 13349, LT2 and ATCC 140285) were used. Bi<strong>of</strong>ilms<br />

were grown on CBD for 24 hours, in Mueller-Hin<strong>to</strong>n Broth, at 37 °C with shaking at<br />

125 rpm. Disinfection was performed with sodium hypochlorite, benzalkonium<br />

chloride, hydrogen peroxide and triclosan, while bacterial death was assessed by<br />

standard plate method on Trypticase Soy Agar after sonication. Results showed<br />

that sodium hypochlorite had the lowest MBEC values while triclosan had the<br />

worst performance, since no S. enterica bi<strong>of</strong>ilm eradication was achieved even at<br />

the maximum concentration used. It was also found that, except for hydrogen peroxide,<br />

MBEC values for L. monocy<strong>to</strong>genes bi<strong>of</strong>ilms were similar or inferior <strong>to</strong> those<br />

found for S. enterica. Significant differences on minimum survival biomass results<br />

were also observed between strains <strong>of</strong> the same species. Summarizing, this work<br />

has pointed out chlorine agents as the most effective on disinfecting bi<strong>of</strong>ilms <strong>of</strong><br />

both species used and revealed a higher susceptibility <strong>of</strong> L. monocy<strong>to</strong>genes bi<strong>of</strong>ilms<br />

<strong>to</strong> disinfection in general when compared with S. enterica bi<strong>of</strong>ilms. Moreover, not<br />

only interspecies but also intraspecies variability were found <strong>to</strong> influence disinfection<br />

efficacy.<br />

Examination <strong>of</strong> the ability <strong>of</strong> ad<strong>here</strong>nce, bi<strong>of</strong>ilm formation and sensitivity<br />

<strong>to</strong> some disinfectants <strong>of</strong> different Listeria monocy<strong>to</strong>genes strains<br />

Milanov, D. 1, Vidić, B. 1, Petrović, J. 1, Bugarski, D. 1 and Ašanin, R. 2<br />

1. Scientific Veterinary Institute “Novi Sad”, Novi Sad, Republic <strong>of</strong> Serbia<br />

2. Faculty <strong>of</strong> Veterinary Medicine, Republic <strong>of</strong> Serbia<br />

The objective <strong>of</strong> this work was <strong>to</strong> examine the capabilities <strong>of</strong> 14 Listeria monocy<strong>to</strong>genes<br />

strains <strong>to</strong> attach on glass and <strong>to</strong> form bi<strong>of</strong>ilm on stainless steel surfaces.<br />

The strains originated from animals (8 strains), food (4 strains) and feed (1 strain).<br />

The referent strain was a human isolate (ATCC 19115). The number <strong>of</strong> L. monocy<strong>to</strong>genes<br />

was determined after 3-hours <strong>of</strong> attachment and after 48-hour incubation<br />

in tryp<strong>to</strong>ne soy broth at 25 °C and 37 °C by the use <strong>of</strong> standard count technique<br />

on blood agar from tenfold dilution. Three days old bi<strong>of</strong>ilm formed on glass surfaces<br />

<strong>of</strong> the selected L. monocy<strong>to</strong>genes strains was treated with paracetic acid and<br />

phenol disinfectants for 5 and 10 minutes. On the stainless steel surface the<br />

bi<strong>of</strong>ilms were formed during 7 days <strong>of</strong> incubation in a tryp<strong>to</strong>ne soy broth supplemented<br />

with 0.6 % yeast extract (TSB-YE) at the temperature <strong>of</strong> 25 °C. The developed<br />

structures were examined using scanning electron microscopy. For all the<br />

examined L. monocy<strong>to</strong>genes strains the number <strong>of</strong> attached bacteria <strong>to</strong> glass slides<br />

for 3 hours <strong>of</strong> incubation in static conditions at 25 °C and 37 °C ranged from 10 2-<br />

10 4 cfu/cm 2. After 48 h <strong>of</strong> incubation the number <strong>of</strong> cells that grow on glass slides<br />

did not depend on initial attachment and for all the strains it was 10 5-10 7 cfu/cm 2.<br />

Among tested Listeria monocy<strong>to</strong>genes strains, significant differences in terms <strong>of</strong><br />

their ability <strong>to</strong> form bi<strong>of</strong>ilm were found. Seven <strong>of</strong> 14 investigated strains <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes did not form bi<strong>of</strong>ilm, and only individual bacterial cells were distributed<br />

over the stainless steel surface. The strains classified as bi<strong>of</strong>ilm producers<br />

formed structures <strong>of</strong> different appearances, from a uniform, confluent monolayer<br />

<strong>of</strong> bacterial cells <strong>to</strong> individual large, three-dimensional cell aggregates. L. monocy<strong>to</strong>genes<br />

cells attached <strong>to</strong> glass surface expressed higher resistance <strong>to</strong> disinfectants<br />

than the cells in suspension.<br />

R E F E R E N C E<br />

D / P<br />

176<br />

R E F E R E N C E<br />

D / P<br />

177


R E F E R E N C E<br />

D / P<br />

187<br />

R E F E R E N C E<br />

D / P<br />

188<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Evolution <strong>of</strong> Listeria monocy<strong>to</strong>genes contamination<br />

in poultry production: from the farms <strong>to</strong> the processing levels<br />

Mansour Chemaly, M., Toquin, M.-T., Courtillon, C., Le Nôtre, Y., Rivoal, K. and Fravalo, P.<br />

AFSSA, France<br />

This investigation is a part <strong>of</strong> a large french program aiming <strong>to</strong> assess and update<br />

data regarding L. monocy<strong>to</strong>genes in poultry flocks (laying hens, broilers, turkeys)<br />

and the impact <strong>of</strong> their introduction in the slaughtering process. Samples were<br />

collected from the flocks (5 bootswabs) which were followed up <strong>to</strong> the slaughterhouses<br />

w<strong>here</strong> sampling consisted <strong>of</strong> swabs from the environment (before and after<br />

cleaning and desinfecting procedures), caeca and products (neck skins and fillets).<br />

The isolation <strong>of</strong> L. monocy<strong>to</strong>genes was performed according <strong>to</strong> the NF EN<br />

ISO11290-1 method and the identification included serotyping. RFLP-PFGE was<br />

carried out in order <strong>to</strong> establish the clonal relationships and <strong>to</strong> trace the contamination<br />

between the farms and the slaughterhouses. 32 % <strong>of</strong> the flocks were tested<br />

positive for Listeria monocy<strong>to</strong>genes. The serotyping <strong>of</strong> L. monocy<strong>to</strong>genes strains<br />

showed that the majority belonged <strong>to</strong> the type 1/2a which presented high genetic<br />

diversity (Simpson Index: 0.95). At the slaughterhouses, the contamination varied<br />

between 4 and 26 % <strong>of</strong> the <strong>to</strong>tal sampling. Positive samples were found in the environment<br />

after cleaning and desinfection and the same isolates were found in the<br />

environment between the slaughtering <strong>of</strong> other batches and on the products.<br />

Strains found at the farm level were clearly different from those found at the<br />

slaughterhouses. Our work highlighted the spreading <strong>of</strong> L. monocy<strong>to</strong>genes in poultry<br />

flocks, residual and cross contamination at the slaughterhouses. Despite a high<br />

level <strong>of</strong> faecal contamination at the farm level, no relation could be established<br />

with the contamination <strong>of</strong> slaughterhouses. This suggests that the primary production<br />

is not the main source <strong>of</strong> contamination <strong>of</strong> poultry products. Efforts<br />

should be focused on the slaughterhouses w<strong>here</strong> inefficient C&D procedures and<br />

cross contamination were the main identified sources <strong>of</strong> contamination.<br />

A regular survey <strong>of</strong> Listeria in ready-<strong>to</strong>-eat foods (2004 – 2009)<br />

Furtado, R., Lore<strong>to</strong> Campos, M., Correia, C., Ferreira, I., Maia, C., Rosa, N., San<strong>to</strong>s, S.,<br />

San<strong>to</strong>s, M. I. and Saraiva, M.<br />

Departamen<strong>to</strong> de Alimentação e Nutrição, Institu<strong>to</strong> Nacional de Saúde Dou<strong>to</strong>r Ricardo Jorge, I.P.,<br />

Lisboa, Portugal<br />

Listeria monocy<strong>to</strong>genes is one <strong>of</strong> the most important pathogens found in food. The<br />

consumption <strong>of</strong> food products contaminated with this bacterium can cause listeriosis,<br />

a disease with a high mortality rate that can affect especially vulnerable<br />

groups. The ubiqui<strong>to</strong>us nature <strong>of</strong> Listeria, its resistance <strong>to</strong> several environmental<br />

conditions and the ability <strong>to</strong> grow at refrigeration temperatures, promote the occurrence<br />

<strong>of</strong> contamination in any stage <strong>of</strong> the food chain. Taking simple precautions<br />

like thoroughly cooking foods, respecting the chill chain, proper washing<br />

fruit and vegetables eaten raw and washing hands repeatedly, reduce contamination<br />

chances. The main activity <strong>of</strong> our Labora<strong>to</strong>ry, it is the control <strong>of</strong> the microbiological<br />

quality <strong>of</strong> ready-<strong>to</strong>-eat foods served in canteens. This regular survey includes<br />

the evaluation <strong>of</strong> the lay out <strong>of</strong> the physical facilities based on Reg. (EU)<br />

n. º 852/2004 and Codex Alimentarius, and the collection <strong>of</strong> food samples and<br />

swabs in surfaces and utensils. This work evaluate the presence <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

(VIDAS LMO2 Bio 12/11-03/04) and quantify the level <strong>of</strong> Listeria spp.<br />

(ISO 11290-2:1998/Amd:2004), in food samples collected in the establishments<br />

surveyed by INSA Lisboa, between 2004 and 2009. In a <strong>to</strong>tal <strong>of</strong> 5450 ready-<strong>to</strong>-eat<br />

foods analysed, Listeria monoy<strong>to</strong>genes was present in 73 samples, in which 8.2 %<br />

exceeded the 100 cfu/g limit. Considering that daily thousands <strong>of</strong> meals are consumed<br />

by risk populations, such as those <strong>of</strong> kindergarten, schools, hospitals and<br />

homes for the elderly, and that our positive results correspond <strong>to</strong> 58.9 % <strong>of</strong> samples<br />

collected in these type <strong>of</strong> establishments, they could stand as a useful <strong>to</strong>ol, demonstrating<br />

that if additional precautions are not implemented, Listeria monocy<strong>to</strong>genes<br />

can be a potential risk for these particularly susceptible groups, even at concentrations<br />

less than 100 cfu/g.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Incidence <strong>of</strong> Listeria monocy<strong>to</strong>genes in Queijo Fresco<br />

Rosa, N., Campos, L., Correia, C., Ferreira, I., Furtado, R., Maia, C., San<strong>to</strong>s, S.,<br />

Cunha, C. I. and San<strong>to</strong>s, M. I.<br />

Departamen<strong>to</strong> de Alimentação e Nutrição, Institu<strong>to</strong> Nacional de Saúde Dou<strong>to</strong>r Ricardo Jorge, I.P.,<br />

Lisboa, Portugal<br />

Listeria monocy<strong>to</strong>genes is a ubiqui<strong>to</strong>us bacterium responsible for inumerous cases<br />

and outbreaks <strong>of</strong> listeriosis in humans, usually transmitted by the consumption<br />

<strong>of</strong> contaminated foods, mainly those “ready-<strong>to</strong>-eat”. Being “Queijo Fresco” a ready<strong>to</strong>-eat<br />

product, it’s one <strong>of</strong> those which have driven Listeria monocy<strong>to</strong>genes <strong>to</strong> be a<br />

major concern <strong>to</strong> Public Health. It was the purpose <strong>of</strong> this study <strong>to</strong> evaluate the<br />

presence <strong>of</strong> Listeria monocy<strong>to</strong>genes in “Queijo Fresco” from a number <strong>of</strong> brands<br />

with the most commercial significance in the Lisbon area. The “Type <strong>of</strong> Commercial<br />

Presentation” was also studied for its influence in the occurrence <strong>of</strong> the microrganism<br />

in the above mentioned food products. A <strong>to</strong>tal <strong>of</strong> 125 samples were examined<br />

for the presence <strong>of</strong> Listeria monocy<strong>to</strong>genes. Secondary enrichments, in<br />

Fraser broth, were analysed by the mini-VIDAS LMO® (bioMérieux, Durham,<br />

France), enzyme-linked fluorescent immunoassay method for Listeria monocy<strong>to</strong>genes<br />

detection. Positive samples were confirmed by isolation on ALOA® (AES<br />

CHEMUNEX) selective agar followed by biochemical characterization with API<br />

LISTERIA® (bioMérieux, Durham, France). Statistical analysis was performed<br />

using the Statistical Package for the Social Sciences (SPSS v13.0) s<strong>of</strong>tware. Of 125<br />

samples, 13 (10.4 %) were positive for Listeria monocy<strong>to</strong>genes. It was statistically<br />

proved that the Commercial Presentation influences the presence <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes, being the cheeses sold packed less contaminated than those sold<br />

unpacked (p = ,049). This study demonstrates that Listeria monocy<strong>to</strong>genes is present<br />

in “Queijo Fresco” from a range <strong>of</strong> brands commercialized in Lisbon. The contamination<br />

observed, represents a potential risk for the Portuguese consumer due<br />

<strong>to</strong> his natural taste for this product.<br />

Risk fac<strong>to</strong>rs for Listeria monocy<strong>to</strong>genes contamination<br />

in French broiler flocks<br />

Aury, K., Le Bouquin, S., Toquin, M.-T., Petetin, I., Le Nôtre, Y., Allain, V.,<br />

Fravalo, P. and Mansour Chemaly, M.<br />

AFSSA, France<br />

Despite a decreasing incidence <strong>of</strong> Listeriosis in France since 1999, Listeria monocy<strong>to</strong>genes<br />

continue <strong>to</strong> be a public health concern. In order <strong>to</strong> update data relating<br />

<strong>to</strong> L. monocy<strong>to</strong>genes in France in poultry production, an epidemiological study was<br />

conducted in broiler chicken flocks between Oc<strong>to</strong>ber 2005 and September 2006<br />

aiming <strong>to</strong> identify the potential risk fac<strong>to</strong>rs associated <strong>to</strong> the presence <strong>of</strong> L. monocy<strong>to</strong>genes.<br />

142 broiler chicken flocks were included in this study. The L. monocy<strong>to</strong>genes<br />

status <strong>of</strong> these flocks was based on 5 samples <strong>of</strong> bootswabs per farm. The<br />

flocks were considered positive if at least one sample was tested positive for L.<br />

monocy<strong>to</strong>genes. Information on potential risk fac<strong>to</strong>rs was collected by questionnaire<br />

at the same time as sample collection. The association between characteristic<br />

management practices and L. monocy<strong>to</strong>genes status was assessed by logistic regression.<br />

The prevalence <strong>of</strong> L. monocy<strong>to</strong>genes contamination in these flocks was<br />

31.7 %. The risk <strong>of</strong> L. monocy<strong>to</strong>genes contamination was increased when farmer<br />

did not respect the principle <strong>of</strong> the two areas (clean and dirty) at the poultry house<br />

entrance. The absence <strong>of</strong> spraying at the first disinfection and/or pest control <strong>of</strong><br />

the poultry house before the arrival <strong>of</strong> the next flock was found <strong>to</strong> increase the<br />

risk <strong>of</strong> being infected. When litter s<strong>to</strong>rage was not protected and when farm staff<br />

take care <strong>of</strong> the other broiler chicken houses <strong>of</strong> the holding, the risk <strong>of</strong> L. monocy<strong>to</strong>genes<br />

contamination increased significantly. For the watering system, pipettes<br />

without recupera<strong>to</strong>r were found <strong>to</strong> be associated with a higher risk <strong>of</strong> contamination<br />

than pipettes with recupera<strong>to</strong>r or drinkers. This study brings new insights<br />

for the risk management <strong>of</strong> L. monocy<strong>to</strong>genes infection in broiler flocks.<br />

R E F E R E N C E<br />

D / P<br />

189<br />

R E F E R E N C E<br />

D / P<br />

190


R E F E R E N C E<br />

D / P<br />

191<br />

R E F E R E N C E<br />

D / P<br />

192<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Portuguese sushi: is it contaminated with L. monocy<strong>to</strong>genes?<br />

Mendes, D., Furtado, R., Maia, C., Correia, C., Campos Cunha, I., Pedroso, L. and San<strong>to</strong>s, M. I.<br />

Institu<strong>to</strong> Nacional de Saúde Dou<strong>to</strong>r Ricardo Jorge<br />

Healthier nutritional lifestyles and cultural globalization have popularized the<br />

consumption <strong>of</strong> ready-<strong>to</strong>-use products <strong>of</strong> seafood, like sushi, that were previously<br />

restricted <strong>to</strong> oriental countries. Sushi is a traditional Japanese food, mostly composed<br />

<strong>of</strong> rice and raw fish. The sushi trend involves a higher consume <strong>of</strong> raw fish,<br />

particularly among young, urban people including pregnant women. Although fish<br />

is considered a healthy food, as with other animal products, consumption <strong>of</strong> raw<br />

muscle incurs potential health risks such as ingestion <strong>of</strong> pathogenic bacteria. Additionally,<br />

the preparation practices involved in the production <strong>of</strong> sushi has the<br />

potential <strong>to</strong> allow contamination, namely with Listeria monocy<strong>to</strong>genes, the agent<br />

<strong>of</strong> listeriosis, an infection that targets mainly pregnant women (and their fetuses),<br />

children, the elderly and immunocompromised individuals. In spite <strong>of</strong> the number<br />

<strong>of</strong> cases per annum is relatively low, these infections can be acute, with mortality<br />

up <strong>to</strong> 30 %. Due <strong>to</strong> the seriousness <strong>of</strong> clinical manifestations and high rates <strong>of</strong> mortality<br />

in populations at risk, control and prevention <strong>of</strong> this disease, represents an<br />

important challenge <strong>to</strong> sanitation authorities and deserves the attention <strong>of</strong> food<br />

microbiologists and health pr<strong>of</strong>essionals. In this study, 90 samples <strong>of</strong> sushi collected<br />

from different establishments in Lisbon, Portugal, were analyzed for their<br />

microbiological status and the prevalence <strong>of</strong> pathogenic bacteria. For L. monocy<strong>to</strong>genes<br />

detection the ISO 11290-2:1998/Amd 1:2004 method was performed. The<br />

results obtained showed that all samples were negative for L. monocy<strong>to</strong>genes and<br />

just one <strong>of</strong> them revealed the presence <strong>of</strong> Listeria seeligeri. Although the number <strong>of</strong><br />

samples studied was small, we believe this study provides important information<br />

concerning the incidence <strong>of</strong> L. monocy<strong>to</strong>genes in Portuguese sushi and also evidences<br />

that the consumption <strong>of</strong> this type <strong>of</strong> food probably is not a major problem<br />

in relation with this pathogen.<br />

Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes throughout the production<br />

process <strong>of</strong> Parma ham: tracing contaminations from slaughterhouses<br />

<strong>to</strong> the final product<br />

Prencipe, V. A. 1, Rizzi, V. 1, Iannetti, L. 1, Serraino, A. 2, Calderone, D. 3, Rossi, A. 4, Morelli, D. 1,<br />

Marino, L. 1 and Migliorati, G. 1<br />

1. Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

2. University <strong>of</strong> Bologna - Veterinary Medicine, Italy<br />

3. Consorzio del Prosciut<strong>to</strong> di Parma, Italy<br />

4. Centro Ricerche Produzioni Animali S.p.A., Italy<br />

In order <strong>to</strong> evaluate Listeria monocy<strong>to</strong>genes prevalence and contamination levels<br />

throughout the Parma Ham production chain, 774 swine carcasses were traced<br />

along the whole process until the final product. Analyses were carried out on samples<br />

originated from the same carcass but taken at three different points <strong>of</strong> the<br />

process (as carcass, fresh ham and dry-cured ham). Fecal samples were also taken<br />

from 498 carcasses (about 2/3 <strong>of</strong> the whole sample). The prevalence <strong>of</strong> Listeria<br />

monocy<strong>to</strong>genes, calculated at 2.9 % in the carcasses swabbed in slaughterhouses,<br />

increased up <strong>to</strong> 12.5 % in fresh hams sampled when entering the manufacturing<br />

plants and eventually fell <strong>to</strong> 2.0 % at the end <strong>of</strong> the production chain (de-boned<br />

and packaged dry-cured hams). Listeria monocy<strong>to</strong>genes was isolated from only one<br />

fecal sample (prevalence 0.2 %), corroborating the evidence <strong>of</strong> the low importance<br />

<strong>of</strong> primary production as a source <strong>of</strong> contamination from this bacteria. Contamination<br />

levels ranged from 0.10 <strong>to</strong> 0.33 MPN/cm 2 in carcasses, 0.04 <strong>to</strong> 2400 UFC/g<br />

in fresh hams, 0.04 <strong>to</strong> 100 UFC/g in dry-cured hams sampled at the end <strong>of</strong> the production<br />

chain. Only the 3.2 % (n = 3) <strong>of</strong> contaminated fresh hams came from contaminated<br />

carcasses, no one among the 14 contaminated dry-cured hams came<br />

from contaminated fresh hams. Certainly, cutting was the stage with the highest<br />

contamination risk throughout the Parma Ham production process. However, beside<br />

the sharp reduction <strong>of</strong> contaminations subsequent <strong>to</strong> dry-curing, seasoning<br />

environment should be addressed as the critical step <strong>of</strong> the whole process for the<br />

contaminations found in the final product. The significant differences between<br />

plants corroborates the importance <strong>of</strong> the processing environment as source <strong>of</strong><br />

contamination from Listeria monocy<strong>to</strong>genes. Further biomolecular analyses are<br />

needed in order <strong>to</strong> confirm the marginal role <strong>of</strong> transferring contaminations between<br />

different stages <strong>of</strong> the production chain and <strong>to</strong> assess the presence <strong>of</strong> persistent<br />

strains inside the processing plants.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Effect <strong>of</strong> the inoculum size on growth <strong>of</strong> L. monocy<strong>to</strong>genes<br />

in dices <strong>of</strong> poultry breast<br />

Lardeux, A.-L., Gnanou-Besse, N., Doux, C. and de Courseulles, E.<br />

AFSSA, France<br />

Traditionally the microbiological safety <strong>of</strong> foods has been established via challenge<br />

tests. Since the contamination level attained by Listeria monocy<strong>to</strong>genes may depend<br />

on the initial bacterial concentration, it is important <strong>to</strong> take in<strong>to</strong> account this effect<br />

when performing challenge tests. To date, t<strong>here</strong> are, <strong>to</strong> our knowledge, few studies<br />

which have examined its impact in food. The effect <strong>of</strong> the inoculum size on growth<br />

<strong>of</strong> L. monocy<strong>to</strong>genes in a matrix <strong>of</strong> dices <strong>of</strong> poultry breast has been studied. This matrix<br />

is interesting because it is located near the interface growth/no growth <strong>of</strong> L.<br />

monocy<strong>to</strong>genes, it has a natural micr<strong>of</strong>lora and contains lactic acid. Challenge tests<br />

have been carried out for three batches, according <strong>to</strong> technical guidance document<br />

<strong>of</strong> Community Reference Labora<strong>to</strong>ry for Listeria monocy<strong>to</strong>genes. Thus, the fastest<br />

strain <strong>of</strong> L. monocy<strong>to</strong>genes has been selected by a study <strong>of</strong> strains growth with<br />

potassium lactate at different concentrations. Then, L. monocy<strong>to</strong>genes growth studies<br />

were carried out at 8 °C, using two contamination levels (1 cfu/g and 100 cfu/g),<br />

in vacuum-sealed bags or under modified atmosp<strong>here</strong> (50 % nitrogen – 50 % carbon<br />

dioxide). Follow-ups <strong>of</strong> natural micr<strong>of</strong>lora (mesophilic and lactic) have been<br />

performed, as well as the measurement <strong>of</strong> physical-chemical parameters (pH, a w<br />

and lactic acid). L. monocy<strong>to</strong>genes was able <strong>to</strong> grow in all conditions. However,<br />

growth was higher under vacuum packaging conditions, and with a high inoculation<br />

level. Moreover, important differences were observed between batches. In parallel,<br />

an important growth <strong>of</strong> background micr<strong>of</strong>lora was observed, which probably had a<br />

strong impact on L. monocy<strong>to</strong>genes growth.<br />

Investigation in<strong>to</strong> the mechanisms <strong>of</strong> detergent induced changes<br />

in disinfectant susceptibility <strong>of</strong> attached Listeria monocy<strong>to</strong>genes<br />

Wal<strong>to</strong>n, J., Hayes, R., Protheroe, R., Hill, D. and Gibson, H.<br />

University <strong>of</strong> Wolverhamp<strong>to</strong>n, UK<br />

The control <strong>of</strong> Listeria through effective cleaning and disinfectant strategies is<br />

crucial <strong>to</strong> maintaining the safety <strong>of</strong> food. The aim <strong>of</strong> this work is <strong>to</strong> investigate the<br />

effect <strong>of</strong> detergent treatments on the susceptibility <strong>of</strong> attached Listeria monocy<strong>to</strong>genes<br />

<strong>to</strong> subsequent disinfectant treatments. While the detergents were prepared<br />

<strong>to</strong> working concentrations, the disinfectants were lower than the recommended in<br />

use concentration <strong>to</strong> allow quantification <strong>of</strong> changes in susceptibility. Results so<br />

far have shown that L. monocy<strong>to</strong>genes, attached <strong>to</strong> stainless steel surfaces, became<br />

significantly less susceptible (up <strong>to</strong> 3.5 log10 difference) <strong>to</strong> benzalkonium chloride<br />

(BAC, 0.005 % v/v) following treatment with the anionic detergents sodium<br />

alkyl sulphate (SAS, 0.2 % v/v), sodium dodecyl sulphate (SDS, 0.2 % w/v) and<br />

sodium lauryl ether sulphate (SLES 0.2 % v/v). L. monocy<strong>to</strong>genes also became significantly<br />

less susceptible (0.4 log10 difference) <strong>to</strong> sodium dichloroisocyanurate<br />

(NaDCC, 0.0008 % w/v) following treatment with SAS but significantly more susceptible<br />

(1 log10 difference) following treatment with SLES. The non-ionic detergent,<br />

fatty alcohol ethoxylate (FAE, 0.1 % v/v), had no effect on susceptibility <strong>to</strong><br />

either BAC or NaDCC. The changes in susceptibility may be due <strong>to</strong> effects on cell<br />

membrane permeability, cell surface hydrophobicity and/or reduced uptake due <strong>to</strong><br />

efflux. Flow cy<strong>to</strong>metry using the fluoresceine propidium iodide revealed significant<br />

increases in cell membrane permeability by SAS and FAE while no change<br />

was observed with SDS. Hydrophobic interaction chroma<strong>to</strong>graphy showed that L.<br />

monocy<strong>to</strong>genes became less hydrophobic following treatment with SAS and SDS<br />

but FAE had no effect. Current work using ethidium bromide has shown a reduction<br />

in fluorescence following treatment with SAS and SDS suggesting that the detergents<br />

may trigger an efflux mechanism resulting in reduced uptake <strong>of</strong> disinfectant.<br />

To conclude, detergents can influence the susceptibility <strong>of</strong> L. monocy<strong>to</strong>genes<br />

<strong>to</strong> BAC and NaDCC which does not appear <strong>to</strong> be related <strong>to</strong> changes in cell membrane<br />

permeability. However, susceptibility does appear <strong>to</strong> correlate with changes<br />

in cell surface hydrophobicity and enhanced efflux.<br />

R E F E R E N C E<br />

D / P<br />

193<br />

R E F E R E N C E<br />

D / P<br />

194


R E F E R E N C E<br />

D / P<br />

195<br />

R E F E R E N C E<br />

D / P<br />

196<br />

AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Prevalence <strong>of</strong> Listeria monocy<strong>to</strong>genes in raw milk sold<br />

at vending machines in Abruzzo region<br />

Prencipe, V. A., Scat<strong>to</strong>lini, S., Sperandii, A. F. and Migliorati, G.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

A survey was carried out in Abruzzo region for assessing the microbial quality <strong>of</strong><br />

raw milk sold at self-service au<strong>to</strong>matic vending machines. From January <strong>to</strong> Oc<strong>to</strong>ber<br />

2008, 299 raw milk samples, collected at 20 distribu<strong>to</strong>rs, were tested <strong>to</strong> detect<br />

Listeria monocy<strong>to</strong>genes. Moreover, the presence <strong>of</strong> the compulsory information<br />

for consumer and the observance <strong>of</strong> the proper s<strong>to</strong>rage temperature were<br />

verified. 31 samples (10.5 %) positive for Listeria monocy<strong>to</strong>genes were found, taken<br />

from 6 distribu<strong>to</strong>rs supplied by 3 farms. 29 over the 31 positive samples were collected<br />

at 4 vending machines supplied by only one farm. All the Listeria monocy<strong>to</strong>genes<br />

strains were identified as serotype 4 b. The resistance <strong>to</strong> two antimicrobials<br />

was detected in 26.6 % <strong>of</strong> the strains tested showing two different patterns<br />

(OXCC, OXL), in 74.2 % <strong>to</strong> three antimicrobials with only one pr<strong>of</strong>ile (OXCCL).<br />

PFGE analysis identified a single pulsotype specific for each farm, whose samples<br />

had been found contaminated by Listeria monocy<strong>to</strong>genes. The combination <strong>of</strong> the<br />

resistance patterns and restriction pr<strong>of</strong>ile <strong>of</strong> the 29 samples isolated from A farm,<br />

identified three different subtypes. The persistence <strong>of</strong> Listeria monocy<strong>to</strong>genes in<br />

raw milk samples, supplied by A farm, confirmed that an inadequate hygiene management<br />

exposes the consumer <strong>to</strong> a real risk <strong>of</strong> infection by Listeria monocy<strong>to</strong>genes.<br />

The absence <strong>of</strong> information for the consumer and the lack <strong>of</strong> cold chain<br />

maintenance during the distribution step could increase this risk. T<strong>here</strong>fore, it is<br />

necessary that farmers and distribu<strong>to</strong>rs’ managers apply continuously the own<br />

check and its efficacy should be verified by the Competent Authority implementing<br />

specific surveillance plans. However it is fundamental <strong>to</strong> inform the consumer<br />

correctly and on continuous basis about the potential risks associated <strong>to</strong> food<br />

products and on the right handling procedures during transport, s<strong>to</strong>rage and treatment<br />

<strong>of</strong> food products.<br />

Characterization <strong>of</strong> Listeria monocy<strong>to</strong>genes strains isolated from s<strong>of</strong>t<br />

and semi s<strong>of</strong>t cheeses sampled at retail level<br />

Acciari, V., Torresi, M., Migliorati, G., Di Giannatale, E., Semprini, P. and Prencipe, V.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

The aim <strong>of</strong> this study was <strong>to</strong> characterize the 47 Listeria monocy<strong>to</strong>genes strains<br />

isolated during a survey carried out on cheeses sampled at retail level. Five semis<strong>of</strong>t<br />

and s<strong>of</strong>t cheeses (gorgonzola, taleggio, asiago, crescenza and brie) were selected<br />

among the higher consumption dairy products in Italy and the most frequently<br />

contaminated by Listeria monocy<strong>to</strong>genes. Each strain was serotyped, tested<br />

for susceptibility <strong>to</strong> antimicrobials and pulsotyped using AscI and ApaI Pulsed<br />

Field Gel Electrophoresis (PFGE). The main serotypes were 1/2a (76.6 %) and 1/2c<br />

(21.3 %), the 1/2b was isolated in only one sample. The antimicrobial resistance<br />

patterns showed that most Listeria monocy<strong>to</strong>genes strains were resistant <strong>to</strong><br />

oxacillin (97.6 %), lincomicin (80.9 %) and clindamycin (78.7 %). The resistance <strong>to</strong><br />

two antimicrobials with two different resistance patterns (OXCC, OXL) was detected<br />

in 17 % <strong>of</strong> the strains tested, <strong>to</strong> three antimicrobials with one resistance<br />

pr<strong>of</strong>ile (OXCCL) in 70.2 %. No strains were sensitive <strong>to</strong> all antimicrobials tested.<br />

Combination <strong>of</strong> AscI and ApaI macrorestriction patterns yielded 11 different pulsotypes<br />

clustering in three groups. Two main pulsotypes were found by grouping<br />

21.3 % and 57.4 % <strong>of</strong> the strains isolated. Evaluation <strong>of</strong> PFGE pr<strong>of</strong>iles showed no relationship<br />

between pulsotypes and cheese type, manufacturer or retail outlet.<br />

Temporal distribution <strong>of</strong> the prevailing pulsotypes showed a persistent pr<strong>of</strong>ile<br />

throughout most <strong>of</strong> the study period, except from August <strong>to</strong> September, when a<br />

different pulsotype was found. T<strong>here</strong>fore the temporal variation in the prevalence<br />

<strong>of</strong> specific strains, as reported in this survey, could be the consequence <strong>of</strong> fac<strong>to</strong>rs<br />

able <strong>to</strong> influence the product contamination pattern. Large scale studies will contribute<br />

<strong>to</strong> assess the dynamics <strong>of</strong> Listeria monocy<strong>to</strong>genes contamination.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 126–177, 187–198<br />

Preliminary report on the organisation <strong>of</strong> a food microbiology<br />

pr<strong>of</strong>iciency testing program as a <strong>to</strong>ol <strong>to</strong> guarantee the equivalence<br />

<strong>of</strong> the US and IT <strong>of</strong>ficial control systems<br />

Di Giannatale, E., Marfoglia, C., Prencipe, V., Salini, R., Migliorati, G. and Ricci, L.<br />

Istitu<strong>to</strong> Zoopr<strong>of</strong>ilattico Sperimentale dell’Abruzzo e del Molise “G. Caporale”, Italy<br />

After Listeria monocy<strong>to</strong>genes was isolated in products imported from Italy, in 2002<br />

the Italian Ministry <strong>of</strong> Health developed an integrated approach <strong>to</strong> assure the<br />

equivalence between the Italian and the US <strong>of</strong>ficial control systems. A key part <strong>of</strong><br />

the ongoing program has been carried out thought an ad hoc labora<strong>to</strong>ry pr<strong>of</strong>iciency<br />

testing with the aim <strong>of</strong> moni<strong>to</strong>ring the technical skills <strong>of</strong> the <strong>of</strong>ficial labora<strong>to</strong>ries in<br />

charge <strong>of</strong> testing Listeria monocy<strong>to</strong>genes on meat products <strong>to</strong> be exported in the<br />

USA. This study describes the organization <strong>of</strong> a pr<strong>of</strong>iciency testing and the pro<strong>to</strong>col<br />

experimented for preparation and verification <strong>of</strong> samples consisting in food matrix<br />

with a basic micr<strong>of</strong>lora contaminated with Listeria monocy<strong>to</strong>genes at two different<br />

level <strong>of</strong> contamination. The pr<strong>of</strong>iciency testing is an indispensable <strong>to</strong>ol <strong>to</strong> assess<br />

the performance level <strong>of</strong> a testing labora<strong>to</strong>ry and <strong>to</strong> demonstrate its reliability at<br />

national and international framework. In our specific context, the pr<strong>of</strong>iciency program<br />

has also concurred as one <strong>of</strong> the effective measures <strong>to</strong> guarantee the equivalence<br />

within the certifications issued by the Italian and US <strong>of</strong>ficial labora<strong>to</strong>ries.<br />

High nisin susceptibility <strong>of</strong> Listeria spp. wild-type strains isolated<br />

from dairies with traditional cheese preservation in Portugal<br />

Pintado, C. M. B. S1,2 and Ferreira M. A. S. S. 2<br />

1. Escola Superior Agrária, Institu<strong>to</strong> Politécnico de Castelo Branco, Castelo Branco, Portugal<br />

2. Institu<strong>to</strong> Superior de Agronomia, Universidade Técnica de Lisboa, Lisboa, Portugal<br />

Evaluation <strong>of</strong> nisin susceptibility <strong>of</strong> 219 Listeria spp. wild-type strains isolated from<br />

milk, cheese and cheese processing environment at given conditions was the aim <strong>of</strong><br />

this work. Minimum inhibi<strong>to</strong>ry concentrations (MIC) were evaluated by the agar<br />

incorporation method, on TSYEGA medium. The variability <strong>of</strong> Listeria spp. susceptibility<br />

was very low especially at pH 5.5, with 98 % <strong>of</strong> isolates showing MIC values<br />

between 10 and 50 IU ml -1 at 37 °C. The increase <strong>of</strong> pH from 5.5 (average <strong>of</strong> cheese<br />

core pH) <strong>to</strong> 6.8 (average <strong>of</strong> cheese rind pH) at 37 °C resulted in an increase in the<br />

MIC values by 5 <strong>to</strong> 20 times (from 10 – 50 <strong>to</strong> 50 – 100 IU) or more (> 100 IU ml 1) in<br />

68 % <strong>of</strong> isolates. An increase in temperature from 20 °C <strong>to</strong> 37 °C at pH 5.5 resulted<br />

in an increase on MIC values only in 8 % <strong>of</strong> all isolates tested. Nisin spontaneous resistance<br />

at a frequency <strong>of</strong> 10 -3 <strong>to</strong> 10 -4 was observed more frequently at pH 6.8 and<br />

37 °C. Exposure <strong>to</strong> combined stress conditions different from the ones tested<br />

<strong>here</strong>in was previously reported <strong>to</strong> increase cell resistance mechanisms and indicates<br />

that changes in virulence may occur concomitantly. Nevertheless, the majority<br />

<strong>of</strong> strains tested were able <strong>to</strong> grow at 10 IU nisin ml -1 but not at 50 IU ml -1 at pH<br />

5.5, showing a very susceptible pr<strong>of</strong>ile. The lower pH (5.5) and temperature (20 °C)<br />

tested, <strong>of</strong>fered better conditions <strong>to</strong> the inhibi<strong>to</strong>ry action <strong>of</strong> nisin which demonstrated<br />

a good potential for use in bioactive coatings as a combined hurdle effect <strong>to</strong><br />

control L. monocy<strong>to</strong>genes strains in dairy products.<br />

R E F E R E N C E<br />

D / P<br />

197<br />

R E F E R E N C E<br />

D / P<br />

198


R E F E R E N C E<br />

E / P<br />

178<br />

R E F E R E N C E<br />

E / P<br />

179<br />

AREA // E // Communication, risk perception and consumer practices – Social sciences in Listeria control<br />

POSTER PRESENTATIONS // P / 178–181<br />

What is an appropriate level <strong>of</strong> protection for Listeria monocy<strong>to</strong>genes<br />

in foodstuffs consumed by vulnerable groups?<br />

Little, C., Gillespie, I., Grant, K., Gormley, F., Mook, P. and McLauchlin, J.<br />

Health Protection Agency Centre for Infections, UK<br />

It is widely recognised that the consumption <strong>of</strong> contaminated food is an important<br />

route <strong>of</strong> transmission <strong>of</strong> listeriosis and a wide range <strong>of</strong> food products have<br />

been shown <strong>to</strong> be associated with both outbreaks and sporadic cases. Food safety<br />

criteria for L. monocy<strong>to</strong>genes in Regulation (EC) No. 2073/2005 are applicable <strong>to</strong><br />

three categories <strong>of</strong> ready-<strong>to</strong>-eat (RTE) foods: absence <strong>of</strong> L. monocy<strong>to</strong>genes in 25 g is<br />

required in RTE foods intended for infants and those for special medical purposes;<br />

while for all other RTE foods L. monocy<strong>to</strong>genes should not exceed 100 cfu/g within<br />

shelf-life. However, vulnerable groups (pregnant women, the immunosuppressed,<br />

the elderly, and many patients in hospitals) are at particular risk <strong>of</strong> infection, and<br />

consumption <strong>of</strong> low levels <strong>of</strong> L. monocy<strong>to</strong>genes may be <strong>of</strong> greater risk when eaten<br />

by these groups. Source attribution <strong>of</strong> L. monocy<strong>to</strong>genes infections in England and<br />

Wales <strong>to</strong> RTE food sources has shown that the most important source for the overall<br />

population were multi-component foods (e.g. sandwiches, pre-packed mixed<br />

salads) (23.1 %). Attribution <strong>of</strong> major sources was similar for the elderly population<br />

(multi-component foods (22.0 %)). However, for pregnant women, beef (12.3 %),<br />

milk/milk products (11.8 %) and fish (11.2 %) were more important sources <strong>of</strong> infection.<br />

Concurrently, a number <strong>of</strong> outbreak investigations in England, Wales and<br />

Northern Ireland have revealed that infections were linked <strong>to</strong> consumption <strong>of</strong><br />

sandwiches served <strong>to</strong> the vulnerable group population in hospital settings. Sandwiches<br />

tested in these investigations contained the outbreak strain <strong>of</strong> L. monocy<strong>to</strong>genes<br />

at very low concentrations (present at < 10/g) and legally would be deemed<br />

safe by the Regulation. However, the legal limit <strong>of</strong> 100 cfu/g is not based upon formal<br />

dose-response formulas and given the recognised rise in listeriosis observed in<br />

the EU, it is questionable whether the legal limit affords the level <strong>of</strong> protection<br />

deemed appropriate <strong>to</strong> protect the vulnerable group population.<br />

ILCD: An Interactive Listeria culture diversity knowledgebase<br />

Ashok Kumar, J. 1, Barbuddhe, S. B. 1, Kalekar, S. 1, Rodrigues, J. 1, Chopade, N. A. 2, Hain, T. 3<br />

and Chakraborty, T. 3<br />

1. ICAR Research Complex for Goa, Ela, Old Goa<br />

2. Department <strong>of</strong> Pathology, Nagpur Veterinary College, Maharashtra Animal and Fishery Sciences<br />

University, Nagpur, India<br />

3. Institute <strong>of</strong> Medical Microbiology, Justus Liebig University, Giessen, Germany<br />

Indian Listeria Culture Database (ILCD) is an online databank <strong>of</strong> pr<strong>of</strong>iles developed<br />

for the Listeria strains isolated in India from various sources. ILCD is based<br />

on a relational database management system that is hyperlinked <strong>to</strong> visualize phenotyping<br />

and genotyping results in the form <strong>of</strong> tables and DNA fingerprint images<br />

for individual strains. The system has been developed using open source LAMP<br />

(Linux, Apache, MySql and PHP) <strong>to</strong>ols. The database contains geographical source<br />

<strong>of</strong> the strain, its lineage, serotype, source <strong>of</strong> isolation (animal/human), year <strong>of</strong> isolation,<br />

phenotypic and genotypic characteristics, antibiotic sensitivity patterns,<br />

and DNA fingerprint (PFGE image). This is an interactive web based database so<br />

that the data can be exchanged between labora<strong>to</strong>ries electronically. A flexible<br />

search system based on systematic comparisons <strong>of</strong> the database entries allows<br />

inter-labora<strong>to</strong>ry comparison <strong>of</strong> pr<strong>of</strong>iles. T<strong>here</strong> are search options, state wise,<br />

serotype wise and so on. The parameters can be selected <strong>to</strong> search the information<br />

required. Whole description <strong>of</strong> a particular strain can be visualized by <strong>click</strong>ing on<br />

identity number <strong>of</strong> the strain. The data can be s<strong>to</strong>red and accessed via internet<br />

browser. The web interface <strong>of</strong> ILCD allows users with no special programming<br />

background or bioinformatics experience <strong>to</strong> examine the results. The user is able<br />

<strong>to</strong> view the description <strong>of</strong> any particular strain e.g. isolated from human or animal<br />

or food source; or retrieve all information about a particular strain. Being the<br />

first <strong>of</strong> its kind, ILCD is expected <strong>to</strong> be a very helpful <strong>to</strong>ol in strengthening the<br />

concept <strong>of</strong> ‘geographic genomics’ and will be very helpful <strong>to</strong> molecular epidemiologists<br />

and those interested in research on Listeria.


AREA // E // Communication, risk perception and consumer practices – Social sciences in Listeria control<br />

POSTER PRESENTATIONS // P / 178–181<br />

Listeria spp. and the domestic environment: consumer knowledge,<br />

attitudes, risk perceptions and food-handling behaviours<br />

Redmond, E. C.<br />

Cardiff School <strong>of</strong> Health Sciences, UK<br />

In Europe, since 2000 incidence <strong>of</strong> listeriosis has increased by >59 %. This increase<br />

has occurred almost exclusively in adults aged > 60 years with listerial bacteremia.<br />

The importance <strong>of</strong> the home as location for acquiring foodborne disease<br />

has prompted numerous studies <strong>to</strong> understanding <strong>of</strong> consumer food safety practices<br />

in the domestic environment. Cognitive and behavioural data is required <strong>to</strong><br />

develop effective communication initiatives <strong>to</strong> improve food-handling practices<br />

and decrease incidence. This study aims <strong>to</strong> review microbiological studies and determine<br />

consumer risk perceptions, knowledge and domestic food-handling behaviours<br />

<strong>of</strong> that may contribute <strong>to</strong> the risk <strong>of</strong> listeriosis. Electronic searches <strong>of</strong><br />

the Internet and library databases, personal-communication with food safety pr<strong>of</strong>essionals<br />

and attendance at international conferences has facilitated the collection<br />

<strong>of</strong> 143 cognitive and behavioural consumer food safety studies and 16 microbiological<br />

surveys <strong>of</strong> the domestic environment carried out over the past 35years.<br />

Listeria has been isolated from up <strong>to</strong> 62 % domestic kitchens and frequently contaminated<br />

items include refrigera<strong>to</strong>rs (2-10 %isolations) and kitchen dishcloths/sinks<br />

(10-84 %isolations). Cumulatively, 8 % <strong>of</strong> all consumer food safety<br />

studies investigated food safety <strong>of</strong> older adults; 4 % <strong>of</strong> pregnant women and 16 %<br />

other consumer groups. Data indicates fewer (17 %) older adults are aware <strong>of</strong> Listeria<br />

compared <strong>to</strong> other consumer groups (32-87 %). Similarly, 83 % adults (aged ><br />

60years) lack knowledge <strong>of</strong> correct refrigera<strong>to</strong>r temperatures, with > 80 % temperatures<br />

exceeding recommendations. Observed risk-related handling malpractices<br />

implemented by adults aged 60-75years are linked <strong>to</strong> pathogenic contamination<br />

<strong>of</strong> kitchen surfaces and ready-<strong>to</strong>-eat (RTE) foods. Considerable<br />

inadequacies in food safety practices and behavioural influences have been identified<br />

that may increase risk <strong>of</strong> listeriosis. Data comparisons between consumer<br />

groups will be presented, including likelihood <strong>of</strong> consuming RTE foods past useby-dates,<br />

frequency <strong>of</strong> checking refrigera<strong>to</strong>r temperatures and cleaning <strong>of</strong> refrigera<strong>to</strong>rs.<br />

Findings will be discussed in the context <strong>of</strong> strategy development and future<br />

consumer food safety communication iniatitives designed <strong>to</strong> decrease<br />

incidence <strong>of</strong> listeriosis.<br />

Do you know the temperature in your refrigera<strong>to</strong>r?<br />

Røssvoll, E. 1, Jacobsen, E. 2, Ueland, Ø. 1, Einar Granum, P. 3 and Langsrud, S. 1<br />

1. Norwegian Institute <strong>of</strong> Food, Fisheries and Aquaculture Research<br />

2. SIFO, National Institute for Consumer Research, Norway<br />

3. Norwegian School <strong>of</strong> Veterinary Science<br />

The use <strong>of</strong> legislation <strong>to</strong> reduce risk is at present used at all steps <strong>of</strong> the food production<br />

chain from the primary producer <strong>to</strong> the s<strong>to</strong>res. It is, however, both difficult<br />

and undesirable <strong>to</strong> manage consumer food handling through legislation. As soon as<br />

the products are bought by the consumer, the knowledge about how the product is<br />

treated further terminates. The aim <strong>of</strong> this study was <strong>to</strong> improve food safety in<br />

the domestic environment by risk analysis <strong>of</strong> consumer food handling and evaluation<br />

<strong>of</strong> risk-reducing measures. We wanted <strong>to</strong> get a better understanding <strong>of</strong> the<br />

interaction between consumer food handling practices and the potential growth <strong>of</strong><br />

pathogens under domestic conditions. To investigate this we conducted a consumer<br />

survey, a case study, and a microbial experiment and showed that 1) consumers<br />

were ignorant <strong>to</strong> or unaware <strong>of</strong> the temperature in their refrigera<strong>to</strong>r, 2)<br />

ready-<strong>to</strong>-eat (RTE) foods were exposed <strong>to</strong> great temperature variations by consumers,<br />

and 3) the fluctuations in temperature resulted in enhanced growth <strong>of</strong><br />

psychrophilic pathogens as Listeria monocy<strong>to</strong>genes. In conclusion, consumers expose<br />

themselves <strong>to</strong> increased risk <strong>of</strong> contracting food borne diseases by being ignorant<br />

<strong>to</strong> or unaware <strong>of</strong> the temperature in their refrigera<strong>to</strong>rs. The temperature<br />

fluctuations found in the case study enhanced pathogenic growth, and can in worst<br />

cases lead <strong>to</strong> outbreaks <strong>of</strong> food borne diseases.<br />

R E F E R E N C E<br />

E / P<br />

180<br />

R E F E R E N C E<br />

E / P<br />

181


R E F E R E N C E<br />

A / P<br />

47<br />

AREA // A // Biology <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATION// A/ 47<br />

Listeria monocy<strong>to</strong>genes agr system: Quorum sensing, or maybe not?<br />

Garmyn, D., Révelin, C. and Piveteau, P.<br />

Université de Bourgogne – INRA, France<br />

Communication, <strong>of</strong>ten referred as Quorum Sensing, is involved in the adaptation<br />

<strong>of</strong> most bacteria <strong>to</strong> their environment. So far, the agr system is the sole communication<br />

system described in the genus Listeria. The agr communication system affects<br />

the biology <strong>of</strong> Listeria monocy<strong>to</strong>genes during saprophytic life (bi<strong>of</strong>ilm formation)<br />

and during infection. We have recently demonstrated that agr expression<br />

was heterogeneous during growth <strong>of</strong> L. monocy<strong>to</strong>genes EGDe as bi<strong>of</strong>ilm. Our observations<br />

did not fit the criteria <strong>of</strong> the Quorum Sensing paradigm. Indeed, Quorum<br />

sensing means that the whole population presents the same phenotype once<br />

the quorum is reached. In order <strong>to</strong> investigate whether the agr system fits the Quorum<br />

Sensing paradigm, or not, we investigated agr expression in situ by combining<br />

gfp reporters, flow cy<strong>to</strong>metry and fluorescent microscopy. This experimental set<br />

up was used <strong>to</strong> revisit au<strong>to</strong>-induction, Quorum Sensing and more recent theories<br />

<strong>of</strong> bacterial communication during the saprophytic life <strong>of</strong> Listeria monocy<strong>to</strong>genes.


AREA // D // Strategies for prevention and control <strong>of</strong> Listeria monocy<strong>to</strong>genes<br />

POSTER PRESENTATIONS // P / 199– 200<br />

Utilization <strong>of</strong> Lac<strong>to</strong>coccus lactis M104, a wild nisin-producing raw milk<br />

isolate, as an antilisterial adjunct in traditional Greek Graviera cheese<br />

processing<br />

Samelis, I., Pappa, E., Bogovic-Matijasic, B. and Rogelj, I.<br />

National Agricultural Research Foundation, Dairy Research Institute, Greece<br />

Recent validation studies have shown that growth inhibition <strong>of</strong> listerial contamination<br />

is assured during processing and s<strong>to</strong>rage <strong>of</strong> traditional Greek Graviera<br />

cheese in compliance with the current E.U. regula<strong>to</strong>ry criteria for a maximum Listeria<br />

monocy<strong>to</strong>genes population <strong>of</strong> 100 cfu/g allowable in ready-<strong>to</strong>-eat foods. Since,<br />

however, the pathogen survived well during ripening, this study aimed at enhancing<br />

L. monocy<strong>to</strong>genes inactivation by using Lac<strong>to</strong>coccus lactis M104, a nisin-producing<br />

(Nis+) raw milk isolate, as an antilisterial adjunct at cheese manufacture. Fresh<br />

Graviera curds with added a commercial starter culture (SC), or the SC plus the<br />

Nis+ strain (SC+M104), were inoculated (3 log cfu/g) with a 3-strain cocktail <strong>of</strong><br />

avirulent L. monocy<strong>to</strong>genes/innocua before molding in a commercial plant. After<br />

brining, cheeses were ripened at 17-18ºC and 90% RH for 20 days and s<strong>to</strong>red at 4ºC<br />

in vacuum. The fate <strong>of</strong> Listeria was moni<strong>to</strong>red during cheese ripening (0, 1, 3, 5, 11,<br />

17, and 23 days) and s<strong>to</strong>rage (40 and 60 days). Changes in lactic acid bacteria (LAB)<br />

populations and main chemical parameters were also moni<strong>to</strong>red, while the Nis+<br />

gene was detected in ripened cheeses and their LAB consortia by PCR. Populations<br />

<strong>of</strong> Listeria indica<strong>to</strong>rs did not grow in the cheeses at any stage; they declined 10-fold<br />

(P < 0.05) within the first 5 days and survived with little death t<strong>here</strong>after. Although<br />

Nis+ colonies and the Nis+ gene were isolated from the SC+M104 cheeses only, no<br />

major differences in Listeria survival occurred between cheeses with or without<br />

the Nis+ strain. All cheeses contained enterocin (A, B and P) plus plantaricin A<br />

genes associated with Enterococcus faecium and Lac<strong>to</strong>bacillus plantarum <strong>of</strong> the<br />

NSLAB flora, and showed similar chemical changes. Thus, conditions prevailing<br />

during Graviera cheese processing might suppress nisin (Nis+) expression by L.<br />

lactis M104 at levels insufficient <strong>to</strong> deliver or enhance Listeria inactivation.<br />

The European Project BASELINE “Selection and improving <strong>of</strong> fit-forpurpose<br />

sampling procedures for specific foods and risks”<br />

Manfreda, G. and De Cesare, A.<br />

Department <strong>of</strong> Food Science – Alma Mater Studiorum University <strong>of</strong> Bologna, Italy<br />

Food Safety Objectives (FSOs) and Performance Objectives (POs) are new criteria<br />

complementing the existing concepts <strong>of</strong> microbiological criteria. To achieve these<br />

objectives it is critically important a harmonisation <strong>of</strong> food safety control procedures.<br />

The EU project “Selection and improving <strong>of</strong> fit for purpose sampling procedures<br />

for specific foods and risks”-BASELINE, funded under the FP7 programme,<br />

intends <strong>to</strong> obtain the following objectives: 1) <strong>to</strong> review the sampling<br />

schemes currently available for food authorities and food producers <strong>to</strong> collect data<br />

for quantitative risk assessment at EU level; 2) <strong>to</strong> assess the relevance and suitable<br />

limit values <strong>of</strong> POs and FSOs for biological risks, including Listeria monocy<strong>to</strong>genes;<br />

3) <strong>to</strong> evaluate the need for new or adapted methods for sampling and testing the<br />

identified biological risks; 4) <strong>to</strong> develop predictive mathematical models for the<br />

target biological risks; 5) <strong>to</strong> validate and harmonise the sampling schemes and the<br />

alternative detection methods developed in the project; 6) <strong>to</strong> share and disseminate<br />

the scientific knowledge coming out from the project <strong>to</strong> stakeholders. The<br />

project results will be translated in clear recommendation <strong>to</strong> the EC as well as end<br />

users and they will have a significant impact on protection <strong>of</strong> human health. A<br />

poster detailing the project objectives, strategies and expected impact will be presented<br />

at the conference.<br />

R E F E R E N C E<br />

D / P<br />

199<br />

R E F E R E N C E<br />

D / P<br />

200


AUTHORS INDEX


Abdalla, S. D/P/128<br />

Abee, T. A/O/05<br />

Acciari, V. C/P/184, D/P/196<br />

Achtman, M. A/P/27<br />

Agyekum, K. B/O/14<br />

Aharonowitz, Y A/O/11<br />

Aish, J. C/P/125<br />

Albano, H. D/P/163<br />

Alessandria, V A/P/32, A/P/38, B/P/64, D/P/164<br />

Allain, V. D/P/190<br />

Allen, V. A/P/29<br />

Almeida, G. A/P/30, A/P/44, C/P/105, C/P/111,<br />

C/P/113, D/P/166, D/P/165<br />

Almeida, M. T. B/P/66<br />

Alonzo, F. B/P/56<br />

Alvarez-Dominguez, C. A/P/00, B/O/19, C/PL/09,<br />

Amagliani, G. C/P/93<br />

Amar, C. C/P/91, C/P/92<br />

Amar, C. F. C/P/85<br />

Ambrosoli, R. D/P/150<br />

Ancora, M. C/P/102<br />

Andersen, J. B. B/P/182<br />

Andrew, P. B/P/54<br />

Andrew, P. D/P/128<br />

Andrew, P. W. A/P/10<br />

Anita, B. A/P/12, A/P/19<br />

Arapakh, S. D/P/172<br />

Armillotta, G. C/P/185<br />

Arnau, J. D/P/136<br />

Arneborg, N. D/P/132<br />

Asakura, H. A/P/22<br />

Ašanin, R. D/P/177<br />

Ashok Kumar, J. C/O/25, E/P/179<br />

Asséré, A. D/P/146<br />

Atanassova, V. D/P/173<br />

Aubry, C. B/O/21<br />

Auchter, M. B/P/50, D/P/148<br />

Aury, K. D/P/190<br />

Aymerich, T. D/O/43, D/P/136<br />

Azeredo, J. C/P/115, D/P/176<br />

Azevedo, I. A/P/30<br />

Bai, F. A/P/36<br />

Barbosa, J. A/P/30, C/P/111, C/P/113, D/P/160,<br />

D/P/163, D/P/165<br />

Barbuddhe, S- B. A/P/06, A/P/07, B/P/49, C/O/25,<br />

C/P/106, C/P/107, C/P/108, C/P/117, D/P/157, E/P/179.<br />

Barile, M. D/P/129<br />

Barre, L. C/P/103<br />

Batista, S. A/O/02<br />

Baysal, A.H. C/P/114<br />

Beaufort, A. D/O/40, D/P/154<br />

Belessi, C. A. D/P/172<br />

Belletti, N. D/O/43<br />

Benjamin, F. D/P/146<br />

Berg, J. D/P/141<br />

Bergis, H. D/O/40, D/P/154<br />

Bergmann, S. B/P/76<br />

Bergstrøm, A. B/P/182<br />

Bernardi, T. D/P/145<br />

Berrang, M. E. D/P/126<br />

Berzins, A. D/P/155<br />

Besse, N. G. C/P/103<br />

Bhosle, S. N. C/P/107, C/P/108<br />

Bianchi, D. M C/P/123<br />

Bielecka, M.K. C/O/24<br />

Bitar, A. P. B/P/58, B/P/59<br />

Blatter, S. D/P/131<br />

Block, C. S. C/O/29<br />

Boelaert, F. C/O/32, D/P/170<br />

Boone, R. C/P/94<br />

Boor, K. J. A/O/13, A/P/31<br />

Borges, S. A/P/30, D/P/163<br />

Borza, A. C/P/119, C/P/120<br />

Boscher, E. A/P/04<br />

Bosley, J. C/P/119<br />

Botello-Morte, L. A/P/43<br />

Bottero, M. T. C/P/124<br />

Boucheix, C. B/O/20<br />

Bover-Cid, S. D/O/43<br />

Bowman, J. P. A/O/03, A/P/17<br />

Boye, M. B/P/182<br />

Brandi, G C/P/93<br />

Braun, E C/O/29<br />

Bremont, S. C/P/90<br />

Brennan, O. A/O/09<br />

Briandet, R. D/P/142<br />

Briers, Y. A/O/07<br />

Brisabois, A. A/P/01, B/P/47, C/P/82, D/P/146<br />

Brisse, S. A/O/06<br />

Bri<strong>to</strong>, L. A/O/02, D/P/153, D/P/167<br />

Bruno, J. C. B/O/22<br />

Bugarski, D. D/P/177<br />

Burall, L. A/P/26<br />

Burgess, S. C. B/P/48<br />

Busia, G. D/P/175<br />

Butler, F. D/O/42<br />

Byrne, B. C/P/119<br />

Cabanes, D. B/O/23, B/P/65, B/P/66, B/P/68, B/P/70<br />

Cabanova, L. D/P/152<br />

Cabo, M.L. A/P/45<br />

Cabrita, P. A/O/02<br />

Caccia, N. D/P/142<br />

Calderone, D. D/P/194<br />

Calendar, R. A/O/10<br />

Calvo, E. A/P/43<br />

Camejo, A. B/O/23, B/P/70<br />

Cammá, C. C/P/102<br />

Campos, L. D/P/189<br />

Campos, M. L. D/P/188<br />

Candeloro, L. C/P/186<br />

Cantinelli, T. A/O/06<br />

Carneiro, L. A/P/44, D/P/166<br />

Caro, V. A/O/06<br />

Carpentier, B. C/P/103<br />

Carranza-Cereceda, C. A/P/00, B/O/19<br />

Carrasco-Marin, E. A/P/00, B/O/19<br />

Carrera, S. A/P/45<br />

Cartinhour, S. A/O/13<br />

Carvalho, F. B/O/23<br />

Carvalho, F. B/P/65<br />

Casey, P. G. B/P/69<br />

Castle, M. E/O/49<br />

Cécile, C. D/P/171<br />

Ceres, C. D/P/129<br />

Ceri, H. D/P/176<br />

Chafsey, I. A/P/08, D/P/142<br />

Chakraborty, T. A/PL/01, B/P/77, B/P/78,<br />

C/O/25, C/P/106, E/P/179<br />

Chambom, C. A/P/08<br />

Chamot, S. D/P/145<br />

Chaturongakul, S. A/P/31<br />

Chaudhuri, S. B/P/71<br />

Chavant, P. D/P/145<br />

Chaves-Lopez, C. D/P/151<br />

Chemaly, M. D/P/147<br />

Chemaly, M. M. C/P/183, D/P/187, D/P/190<br />

Chen, J. A/O/04, A/P/36<br />

Chen, Y. A/P/16<br />

Chenal-Francisque, V. A/O/06, C/P/100, D/O/46<br />

Chiarini, E. D/P/134<br />

Chopade, N. A. E/P/179<br />

Christensen, B. B. B/P/182<br />

Civera, T. C/P/123, C/P/124<br />

Clark, A. D/O/38<br />

Cocolin, L. A/P/32, A/P/38, D/P/150, D/P/164<br />

Colaneri, C. A/P/01<br />

Comaposada, J. D/P/136<br />

Commeau, N. D/P/154<br />

Corbett, D. B/P/63<br />

Corde, Y. B/P/47<br />

Cormica, M. G. C/P/110<br />

Cornu, M. D/O/40, D/P/154<br />

Correia, C. D/P/188, D/P/189, D/P/191<br />

Cortesi, M. L. D/P/129<br />

Cossart, P. B/O/20<br />

Costas, B. D/O/37<br />

Courvalin, P. A/P/01, C/P/90<br />

Courtillon, C. C/P/183, D/P/187


Couture, H. C/P/97<br />

Crerar, S. E/O/49<br />

Cross, L. C/P/92<br />

Cummins, J. B/P/62<br />

Cunha, C. I. D/P/189, D/P/191<br />

Czuprynski, C. B/P/60<br />

D’Orazio, V. A/P/43<br />

da Silva Felício, M.T C/O/32<br />

da Silva, M. B/O/16<br />

Dal Bello, B. D/P/150<br />

Dalgaard, P. D/O44<br />

Dalmasso, A. C/P/124<br />

D’Amico, D. C/P/112<br />

Daneshvar, M. I. A/P/33<br />

Daniel, P. D/P/171<br />

Danielsson-Tham, M.-L. C/O/30, C/P/122<br />

Datta, A. A/P/26, B/P/67<br />

David, C. B/P/54<br />

Davis, A. D/P/159<br />

D’Costa, D. C/P/106, C/P/108<br />

De Cesare, A. C/P/116<br />

de Courseulles, E. D/P/192<br />

De Martinis, E. C. D/P/148<br />

De Valk, H. C/O/31, C/P/100<br />

Dear, P. C/P/85<br />

Decastelli, L. C/P/123<br />

DeLappe, N. C/P/110<br />

Dell’Era, S. A/O/07<br />

Den Bakker, H.C. A/P/33, D/P/144<br />

Deng, X. A/O/08<br />

Desai, M. C/P/92<br />

Deschamps, J. D/P/142<br />

Destro, M. T. D/P/134, D/P/135<br />

Desvaux, M. A/P/08, D/P/142, D/P/145<br />

Desvaux, M. D/P/142<br />

Dhuri, R. B. C/P/108<br />

di Febo, T. C/P/185<br />

di Giannatale, E. C/P/186, D/P/196, D/P/197<br />

di Pasquale, F. C/P/101, D/P/151<br />

Disson, O. C/O/24<br />

Doijad, S. C/P/107, C/P/108, D/P/157<br />

Dolci, P. D/P/164<br />

Donaldson, J. R. B/P/48<br />

Donnelly, C. C/P/112<br />

Dorey, M. C/P/119, C/P/120<br />

Douey, D. C/P/119, C/P/120<br />

Doux, C. D/P/192<br />

Dowd, G. B/P/69<br />

Duarte Cruz, C. B/P/53<br />

Dumas, E. A/P/08<br />

Dusch, V. D/O/46, C/P/100<br />

Dussurget, O. B/P/55<br />

Ebersbach, T. B/P/182<br />

Egorova, I. C/P/95, C/P/96<br />

Einar Granum, P. E/P/181<br />

Eiriz, E. A/P/45<br />

Eisenreich, W. B/O/18<br />

Elemam M. M. C/P/88<br />

Eleni, S. A/P/12<br />

Eliav, H. C/O/29<br />

Ells, T. C. A/P/40<br />

Eloranta, K. C/P/120<br />

Endres, J. D/P/148<br />

Engeljohn, D. D/PL/11<br />

Ermolaeva, S. A/P/34, B/O/15, B/P/73<br />

Eugster, M. R. A/O/10, A/P/28<br />

Eusébio, C. A/P/44<br />

Eylert, E. B/O/18<br />

Fablet, A. D/P/147<br />

Faith, N. B/P/60<br />

Faith, N. G. B/P/74<br />

Faleiro, M. L. A/P/10<br />

Fall, P. A. D/P/137<br />

Fang, W. A/O/04, A/P/36<br />

Farber, J. A/P/09, C/PL/08, C/P/97, D/P/143<br />

Fávero, L.M. D/P/139<br />

Favretti, M. C/P/84<br />

Felício, T. D/P/170<br />

Félix, B. C/P/82<br />

Fernandes, L. P. D/P/140<br />

Fernandez Guerrero, M. L. B/P/75<br />

Fernandez-Prie<strong>to</strong>, L. A/P/00, B/O/19<br />

Ferreira, I. D/P/188, D/P/189<br />

Ferreira, M.A.S.S. A/P/42, D/P/198<br />

Ferreira, P. B/O/23<br />

Ferreira, R.B. A/O/02<br />

Ferreira, V. A/P/30, D/P/160<br />

Ferreira-Dias, S. D/P/153<br />

Ferrini, G. D/P/136<br />

Fertikov, V. C/P/96<br />

Filiatrault, M. J. A/O/13<br />

Fletcher, G. B/P/53<br />

Flieger, A. A/P/03<br />

Foglini, M. C/P/93<br />

Forster, B. M. B/P/58<br />

Fortes, E. D. A/P/33, D/P/144<br />

Foster, T. D/P/159<br />

Fox, E. D/P/130<br />

Fraccalvieri, R. A/P/37<br />

Franco, B. D. G. M. C/P/118, D/P/134,<br />

D/P/135, D/P/139<br />

Frank, J. D/P/126<br />

Fravalo, P. C/P/183, D/P/187, D/P/190<br />

Freitag, N. B/P/56<br />

Freitag, N. E. B/O/22<br />

Frewer, L.J. E/PL/14<br />

Furtado, D. N. D/P/135<br />

Furtado, R. D/P/188, D/P/189, D/P/191<br />

Fuss, A. A/O/11<br />

Gaedt Kastbjerg, V. A/P/23, B/P/72<br />

Gahan, C.G. B/P/61, B/P/62, B/P/69<br />

Galander, S. A/P/03<br />

Gallagher, D. C/O/28<br />

García-del Portillo, F. A/P/43<br />

Garg, S. D/P/157<br />

Garland, C. D. D/O/38<br />

Garriga, M. D/O/43, D/P/136<br />

Garry, P. D/P/154<br />

Garvey, P. C/P/110<br />

Gendel, S. C/P/97<br />

George, C. D/P/171<br />

Gião, M. S. D/P/127<br />

Gibbs, P. C/P/115, D/P/160<br />

Gibson, H. D/P/193<br />

Gillespie, I. C/O/26, C/O/27, C/P/91, E/P/178<br />

Gillespie, I. A. C/O/33,C/P/98<br />

Gilmartin, N. C/P/99<br />

Gilmour, M. A/P/29, C/PL/08<br />

Glenn, S. B/P/54, D/P/128<br />

Gloria, L. C/P/122<br />

Gnanou-Besse, N. D/P/192<br />

Goebel, W. B/O/18<br />

Gopal, S. A/O/11<br />

Górgolas, M. B/P/75<br />

Gormley, F. E/P/178<br />

Gotz, A. B/O/18<br />

Gouin, E. B/O/20<br />

Goulão, M. M. A/P/42<br />

Goulet, V. C/O/31, D/O/46, C/P/100<br />

Gounadaki, A. S. D/P/172<br />

Graham, S. A/P/29<br />

Gram, L. A/P/23, B/P/72<br />

Granier, S. A. A/P/01<br />

Grant, K. C/O/26, C/O/27<br />

Grant, K. C/O/33, C/P/85, C/P/91, C/P/92,<br />

C/P/98, E/P/178<br />

Grassi, M. A. C/P/123<br />

Graves, L. C/O/36, A/P/33<br />

Grépinet, O. B/P/47<br />

Groelly, J. D/P/145<br />

Guevara, R. E. C/P/86, C/P/89, E/O/47<br />

Guillet, C. C/O/24<br />

Gunjal, P. A/P/06<br />

Haahr Kallipolitis, B. A/P/24<br />

Haase, J. A/P/27<br />

Hagen, N. A/O/11<br />

Hain, T. B/P/77, B/P/78<br />

Hain, T. E/P/179


Halbedel, S. A/P/03<br />

Halberg Larsen, M. A/P/15<br />

Handan Baysal, A D/P/162<br />

Hara, K. C/P/119<br />

Hayes, R. D/P/193<br />

Hearty, S. C/P/99<br />

Heavin, S. A/O/09, A/P/13<br />

Hébert, K. D/P/143<br />

Hebraud, M. D/P/145<br />

Hébraud, M. A/P/08, D/P/142<br />

Heir, E. D/P/161<br />

Helsel, L. O. A/P/33<br />

Henriques, A. B/P/65<br />

Herrera, J.J.R. A/P/45<br />

Hershko-Klement, A. C/O/29<br />

Herve-Bazin, M. C/P/90<br />

Higuchi, Y. A/P/35<br />

Hill, C. B/PL/06, B/P/50, B/P/69<br />

Hill, D. D/P/193<br />

H<strong>of</strong>, H. A/P/27<br />

Hogg, T. C/P/105, C/P/113, D/P/160, D/P/165<br />

Holch, A. B/P/72<br />

Holck, A. D/P/141<br />

Holman, D. B. A/P/40<br />

Hormazábal, J.C. D/PL/12<br />

Hunt, K. D/P/130<br />

Huszczynski, G. C/P/119, C/P/120<br />

Huwiler, S. A/P/28<br />

Hyytia-Trees, E. C/O/36<br />

Iannetti, L. D/P/194<br />

Ida, M. A/P/35<br />

Igimi, S. A/P/18, A/P/22<br />

Imh<strong>of</strong>, R. D/O/41<br />

Ingmer, H. A/P/15, A/P/23, D/P/132<br />

Ioannis, D. A/P/12, A/P/19<br />

Iugovaz, I. C/P/94<br />

Jablonska, J. B/P/81<br />

Jacobsen, E. E/P/181<br />

Jakabi, M. C/P/118<br />

Javanmardi, F. D/P/133<br />

Jean, A. C/P/97<br />

Jen, C. B/P/54<br />

Jenö, P. A/O/02<br />

Jiang, L. A/P/36<br />

J<strong>of</strong>ré, A. D/P/136<br />

Johansson, J. A/O/12, A/P/46<br />

Johnson, D. D/P/156<br />

Join-Lambert, O. C/O/24<br />

Jordan, K. D/O/42, D/P/130, D/P/172<br />

Joyce, S. A. B/P/61, B/P/69<br />

Jusdado, J. J. B/P/75<br />

Kai, A. A/P/35<br />

Kalebar, S. C/P/106, C/P/107, C/P/108, E/P/179<br />

Kallipolitis, B. A/P/02<br />

Kalorey, D. R. C/P/106<br />

Kalorey, D. R. A/P/06, A/P/07, B/P/49,<br />

C/O/25, C/P/117<br />

Kaminskaya, A. A/P/34<br />

Kanarata, S. C/P/109<br />

Kaneko, S. A/P/35<br />

Kantikova, M. D/P/152<br />

Karatzas, K. A/O/09<br />

Kathariou, S. B/P/74<br />

Kawamo<strong>to</strong>, K. B/O/16, C/P/121<br />

Kearney, A. A/P/29<br />

Keevil, C. W. D/P/127<br />

Keich, U. A/O/13<br />

Kent, H. A/P/29<br />

Kerouan<strong>to</strong>n, A. B/P/47<br />

Kiil Nielsen, P. A/P/02, A/P/24<br />

Kim, J. B/P/74<br />

King, L. C/O/31, C/P/100<br />

Klumpp, J. A/O/01<br />

Knabel, S. J. C/O/35, C/P/124<br />

Knoll, L. B/P/60<br />

Knudsen, G. B/P/182<br />

Kolbasov, D. C/P/95, C/P/96<br />

Korikanthimath, V. S. C/O/25<br />

Korkeala H. A/P/11, D/P/155<br />

Koshy, A. A/O/03<br />

Kotwal, S. C/P/117<br />

Koutsoumanis, K. D/O/37<br />

Kreft, J. A/O/11<br />

Krishnendu, M. B/P/77<br />

Kurkure, N. V. A/P/07, B/P/49, C/O/25<br />

Kusumo<strong>to</strong>, A. C/P/121<br />

Lad, M. D/P/159<br />

Ladely, S. D/P/126<br />

Laksanalamai, P. A/P/26<br />

Laksanalamai, P. B/P/67<br />

Landgraf,M. D/P/135<br />

Langsrud, S. D/P/161, E/P/181<br />

Lardeux, A,-L. D/O/40, D/P/192<br />

Larios, O. A/P/29<br />

Larsen, M. H. A/P/23, D/P/168<br />

Larsson, J.T. C/O/34<br />

La<strong>to</strong>rre, L. A/P/37, C/P/87, C/P/116<br />

Laurent, E. C/P/100, D/O/46<br />

Lawrence, M. L. B/P/48<br />

Le Bouquin, S. D/P/190<br />

Le Monnier, A. C/O/24, C/P/90<br />

Le Nôtre, Y. C/P/183, D/P/187, D/P/190<br />

Leclercq, A. A/O/06, C/O/24, C/P/90,<br />

C/P/100, D/O/46<br />

Lecuit, M. A/O/06, B/PL/03, C/O/24, C/P/83,<br />

C/P/90, D/O/46<br />

Lee Chang, K. J. A/O/03<br />

Lee, B. A/P/29<br />

Leisner, J.J. A/P/15<br />

Leite, I. F. C/P/115<br />

Lelli, R. C/P/185<br />

Lemos Vaz-Velho, M D/P/169<br />

Lengeling, A. B/P/76<br />

Li, Z. A/O/08<br />

Licht, T. R. B/P/182<br />

Lienenklaus, S. B/P/81<br />

Lindbäck, T. B/P/79<br />

Lindström, M. A/P/11<br />

Little, C. C/P/91, E/P/178<br />

Little, C. L. C/O/33, C/P/98<br />

Liu, J. C/P/119<br />

Lobacz, A. D/O/45<br />

Loessner, M. J. A/O/01, A/O/07, A/O/10, A/P/28<br />

Lombard, B. D/P/146<br />

Lomonaco, S. C/P/123, C/P/124<br />

Loose, M. B/P/78<br />

Lopes, J.F. C/P/115<br />

Lopez, J. A/O/06<br />

Lopez-Valladares, G. C/O/30<br />

Lortholary, O. C/O/24<br />

Lourenço, A. A. D/P/167<br />

Louro, P. D/P/153<br />

Luber, P. C/PL/10<br />

Luchansky, J. B. B/P/74<br />

Luciani, M. C/P/185<br />

MacDonald, D. C/PL/08<br />

Machado, H. D/P/167<br />

Maciag, P. C. B/O/17<br />

Madarame, H. B/P/52<br />

Madrazo-Toca, F. A/P/00, B/O/19<br />

Magalhães, R. A/P/30, A/P/44, C/P/105,<br />

C/P/111, C/P/113, D/P/166, D/P/165<br />

Magnani, M. C/P/93<br />

Mahmoodi,S. M. M. D/P/133<br />

Maia, C. D/P/188, D/P/189, D/P/191<br />

Maia, C. H. A/P/42<br />

Maia, R. L. E/O/48<br />

Mailles, A. C/P/83<br />

Makela, P. C/O/32<br />

Makela, P. D/P/170<br />

Makino, S.I . B/O/16,C/P/121<br />

Maklon, K. C/P/121<br />

Malik, S.V. S. C/O/25, C/P/106<br />

Mamzer-Bruneel, M.F. C/O/24<br />

Manfreda, G. C/P/116<br />

Marault, M. C/P/82<br />

Marceddu, M. D/P/174<br />

Marfoglia, C. C/P/184, D/P/197


Marie-Christin, M. D/P/171<br />

Marino, L. D/P/194<br />

Mariscotti, J. A/P/43<br />

Marques, N. A/P/10<br />

Marquis, H. B/P/58, B/P/59<br />

Martins, C. G. D/P/139, C/P/118<br />

Martins, M. B/P/68<br />

Maruyama, T. A/P/22<br />

Mascola, L. C/P/86, C/P/89<br />

Mastronicolis, S. A/P/12, A/P/19<br />

Mateus, T. E/O/48<br />

Ma<strong>to</strong>s, J. C/P/91, C/P/92<br />

Matsubara, S. B/O/16<br />

Mattila, M. A/P/11<br />

Mazza, R. D/P/174<br />

Mazzette, R. D/P/174, D/P/175<br />

McCarthy, S. D/P/156<br />

McIlwham, S. A/P/09<br />

McKeown, P. C/P/110<br />

McLauchlin, J. C/P/98, E/P/178<br />

Meinersmann, R. D/P/126<br />

Mejlholm, O. D/O/44<br />

Mele, P. C/P/87<br />

Meloni, D. D/P/174, D/P/175<br />

Mena, C. D/P/166<br />

Mendes, D. D/P/191<br />

Mereghetti, L. B/P/47<br />

Miccolupo, A. A/P/37<br />

Migliorati, G. C/P/102, C/P/184, C/P/186,<br />

D/P/194, D/P/195, D/P/196, D/P/197,<br />

Milanov, D. D/P/177<br />

Milillo, S. R. A/P/33<br />

Milohanic, E. A/P/20<br />

Moes, S. A/O/02<br />

Moller Nielsen, E. C/O/34<br />

Monden, S. A/P/18, A/P/22<br />

Monk, I. R. B/P/50, B/P/51<br />

Mook, P. C/O/26, C/O/27, C/O/33, C/P/91,<br />

C/P/98, E/P/178<br />

Morax, L. A/P/28<br />

Morelli, D. D/P/194<br />

Morey, R. E. A/P/33<br />

Mormile, A. D/P/129<br />

Morrissey, J. A/P/13<br />

Morvan A. C/P/90<br />

Mosher, M. C/P/119, C/P/120<br />

Moubareck, A. C/P/90<br />

Moubarek, C. A/P/01<br />

Moussan Ake, F. A/P/21<br />

Mraheil, M. A. B/P/77<br />

Müller, S. A/P/25<br />

Mureddu, A. D/P/174, D/P/175<br />

Murray, C. C/P/97<br />

Murru, N. D/P/129<br />

Mutanda, C. C/P/119<br />

Nadon, C. A/P/29, C/PL/08<br />

Nakama, A. A/P/22, A/P/35<br />

Nanduri, B. B/P/48<br />

Netterling, S. A/P/46<br />

Neuhaus, K. A/P/25<br />

Neves, E. D/P/153<br />

Nightingale, K. A/P/16<br />

Nijthavorn, N. C/P/109<br />

Nilsson, R. E. A/P/17<br />

Nir-Paz, R. C/O/29<br />

Normanno, G. A/P/37<br />

Noronha, L. D/P/158<br />

Nucera, D. M. C/P/123<br />

Nwaiwu, O. A/P/39, D/P/159<br />

O’ Kennedy, R. C/P/99<br />

O’Brien, M. D/P/130<br />

O’Brien, S. J. C/O/26<br />

O’Byrne, C. A/O/09, A/P/13,A/P/41<br />

Oevermann, A. B/P/52<br />

Okada, N. A/P/18<br />

Okada, Y. A/P/18, A/P/22<br />

Okutani, A. A/P/22<br />

Oliveira, M. A. D/P/148<br />

Oliveira, R. D/P/176<br />

Oliveira, W. P. D/P/140<br />

Oliver, H.F. A/O/13<br />

Omiccioli, E. C/P/93<br />

Ondrusch, N. A/O/11<br />

Orsi, R. H. A/P/33, A/O/13<br />

Padilha Silva, W. A/P/05<br />

Pagot<strong>to</strong>, F. A/P/09, C/PL/08, C/P/94, D/P/143<br />

Paitan, Y. C/O/29<br />

Paparella, A. C/P/101, D/P/151<br />

Parihar, V.S. C/P/117<br />

Parisi, A. A/P/37, C/P/87, C/P/116<br />

Pasche, B. B/P/76<br />

Paterson, Y. B/O/17<br />

Patil, V. A/P/07<br />

Paz, R.-N. A/O/10<br />

Pedroso, L. D/P/191<br />

Pepe, O. D/P/129<br />

Pereira de Martinis, E. C. D/P/140<br />

Petetin, I. D/P/190<br />

Petrovic, J. D/P/177<br />

Phillippy, A.M. A/O/08<br />

Pihier, N. C/P/100, D/O/46<br />

Pillich, H. B/P/78<br />

Pinfold, T. A/O/03<br />

Pintado, C.M.B.S. A/P/42, D/P/198<br />

Pin<strong>to</strong>, E. A/P/10<br />

Piras, F. D/P/174, D/P/175<br />

Pittman, J.R. B/P/48<br />

Pizarro-Cerdá, J. B/O/20<br />

Plaza, B. M. B/P/75<br />

Pompilii, C. C/P/185, C/P/186<br />

Ponnala, L. A/O/13<br />

Portanti, O. C/P/185<br />

Portnoy, D. B/PL/04<br />

Pouillot, R. C/O/28, C/P/97<br />

Poulsen, K. P. B/PL/05, B/P/60<br />

Preising, N. P. B/P/50<br />

Prencipi, V. C/P/102, D/P/196, D/P/197<br />

Prencipe, V. A. C/P/184, D/P/194, D/P/195<br />

Principe, V. C/P/185<br />

Protais, J. D/P/147<br />

Protheroe, R. D/P/193<br />

Pucciarelli, M.G. A/P/43<br />

Pushkareva, V. B/O/15<br />

Quinon, E. D/P/145<br />

Raengpradub, S. A/P/31<br />

Ramos-Vivas, J. A/P/00<br />

Rantsiou, K. A/P/32, A/P/38, D/P/150, D/P/164<br />

Rawool, D. B. C/O/25<br />

Redmond, E. C. E/P/180<br />

Rees, C. D/P/159<br />

Rees, C.E.D. A/P/39<br />

Reis, F. B. D/P/140<br />

Renier, S. D/P/142<br />

Ribot, E. C/O/36<br />

Ricci, L. C/P/186, D/P/197<br />

Riedel, C. U. B/P/51, B/P/50, D/P/148<br />

Ris<strong>to</strong>ri, C.A. D/P/139, C/P/118<br />

Rivoal, K. D/P/147, D/P/187<br />

Rizzi, V. C/O/32, C/P/102, C/P/184, D/P/194<br />

Roberts, I. S. B/P/54, B/P/63<br />

Roche, S. M. B/P/47<br />

Rodrigues, D. D/P/176<br />

Rodrigues, J. A/P/06, C/P/106,C/P/107,<br />

C/P/108, E/P/179<br />

Rodriguez-Del Rio, E. A/P/00, B/O/19<br />

Ronholm, J. A/P/14<br />

Ro<strong>of</strong>, S. E. A/P/33<br />

Rørvik, L. M. B/P/79<br />

Rosa, N. M. C/P/87<br />

Rosa, N., D/P/188, D/P/189<br />

Ross, T. A/P/17<br />

Ross, W. C/P/97<br />

Ross, W. H. A/P/16<br />

Rosshaug, P. S. D/P/168<br />

Rossi, A. D/P/194<br />

Røssvoll, E. E/P/181<br />

Roussel, S. A/P/01, C/O/34,C/P/82<br />

Rousselm, S. D/P/146


Rowlands, R. E. G. C/P/118, D/P/139<br />

Rubinstein, E. B/O/20<br />

Saá Ibusquiza, P. A/P/45<br />

Sabol, A. C/O/36<br />

Sahaghian, R. B/P/74<br />

Sahu, S. B/P/67<br />

Salah, S. D/O/46<br />

Salem, S. C/P/100<br />

Salini, R. D/P/197<br />

Salvat, G. D/P/147<br />

Salzberg, S. L. A/O/08<br />

Santagada, G. A/P/37<br />

San<strong>to</strong>s, I. A/P/30, A/P/44, C/P/111, C/P/113,<br />

D/P/166, D/P/165<br />

San<strong>to</strong>s, M. I. A/P/42, D/P/188, D/P/189, D/P/191<br />

San<strong>to</strong>s, S. D/P/188, D/P/189<br />

Saraiva, M. D/P/188<br />

Sauders, B.D. A/P/33<br />

Savelli, C. C/PL/08<br />

Scat<strong>to</strong>lini,S. D/P/195<br />

Schechner, V. C/O/29<br />

Sc<strong>here</strong>r, S. A/P/25<br />

Schirmer, B. C. D/P/141<br />

Schittler, L. D/P/138<br />

Schlech, W. B/P/80<br />

Schlech, W.F. C/PL/07<br />

Schughart, K. B/P/76<br />

Schuler, S. B/P/54<br />

Schuppler, M. A/O/07<br />

Schvartzman, S. D/O/42, D/P/172<br />

Scortti, M. C/O/24<br />

Scott, V. N. A/P/16<br />

Seavey, M. M. B/O/17<br />

Secic, I. B/P/79<br />

Selyaninov, Y. C/P/95<br />

Semprini, P. C/P/185, C/P/186, D/P/196<br />

Serio, A. C/P/101, D/P/151<br />

Serraino, A. D/P/194<br />

Seuberlich, T. B/P/52<br />

Sewell, D. B/O/17<br />

Shama, G. D/P/128<br />

Shaw, I. A/P/41<br />

Shimojima, Y. A/P/35<br />

Silins I. D/P/155<br />

Silva, A. D/P/153<br />

Silva, J. A/P/44, D/P/158, D/P/163<br />

Silva, W. P. D/P/138<br />

Silveira Naleiro, E. A/P/05<br />

Simpson, J. M. A/P/16<br />

Skandamis, P. N. D/P/172<br />

Skovager, A. D/P/132<br />

Škun<strong>to</strong>va, O. D/P/152<br />

Slepkov, E. R. B/P/58, B/P/59<br />

Smith, M. A. B/O/14<br />

Solodova, E. B/P/81<br />

Somervuo, P. A/P/11<br />

Somov, G.P. B/P/73<br />

Sonegaonkar, A. A/P/07<br />

Sorvillo, F. C/P/86<br />

Sousa, S. B/O/23, B/P/66, B/P/68<br />

Souza, V. M. D/P/140<br />

Sperandii, A. C/P/102<br />

Sperandii, A. F. D/P/195<br />

Stabler, R. A/P/05<br />

Stahl, J.-P. C/P/83<br />

Steigerwalt, A. G. A/P/33<br />

Stephan, R. A/O/01, D/P/131<br />

Stessl, B. C/P/104<br />

S<strong>to</strong>ll, R. A/O/11, B/O/18<br />

S<strong>to</strong>llewerk, K. D/P/136<br />

Sukhapesnaa, J. C/P/109<br />

Sun, Q. A/O/13<br />

Suzuki, H. A/P/18, A/P/22<br />

Swain, B. K. C/O/25<br />

Swaminathan, B. A/P/33<br />

Switt, A. M. D/P/144, D/P/160<br />

Tam Dao, T. C/P/82, D/P/146<br />

Tamburro, M. C/P/85<br />

Tasara, T. A/O/01, D/P/131<br />

Teixeira, A. P. B/P/47<br />

Teixeira, P. A/P/30, A/P/44, C/P/105, C/P/111,<br />

C/P/113, C/P/115, D/P/158, D/P/160, D/P/163,<br />

D/P/165, D/P/166, E/O/48<br />

Teixeira, P. D/P/176<br />

Témoin, S. B/P/47<br />

Tenenhaus-Aziza, F. D/O/42<br />

Tham, M. D. C/P/117<br />

Tham, T. N. B/O/20<br />

Tham, W. C/O/30, C/P/117<br />

Thierry-Bled, F. D/O/46<br />

Thomaz, M. R. S. D/P/140<br />

Todorov, S. D/P/134, D/P/135, D/P/169<br />

Tola, S. C/P/87<br />

Tonucci, F. C/P/93<br />

Toquin, M. -T. C/P/ , D/P/ , D/P/<br />

Torres, R. B/P/75<br />

Torresi, M. C/P/184, D/P/196<br />

Tor<strong>to</strong>rello, M. L. A/O/08<br />

Tran, C. A/O/06<br />

Tresse, O. D/O/39<br />

Trigo, M.J. A/O/02<br />

Trottier, Y. C/P/94<br />

Trout-Yakel, K. A/P/29<br />

Truelstrup Hansen, L. A/P/40<br />

Tyler, S. A/P/29<br />

Ueland, Ø. E/P/181<br />

Umap, S. A/P/07<br />

Upham, J. C/P/119, C/P/120<br />

Utratna, M. A/P/41<br />

Vaikousi, H. D/O/37<br />

Vaillant, V. C/O/31, C/P/83, C/P/100<br />

Valinsky, L. C/O/29<br />

van der Veen, S. A/O/05<br />

Van Domselaar, G. A/P/29<br />

Van Stelten, A. A/P/16<br />

Vandevelde, M. B/P/52<br />

Varfolomeev, A. A/P/34<br />

Vasantrao Kurkure, N. A/P/06<br />

Vázquez, D. A/P/45<br />

Vazquez-Boland, J. C/O/24<br />

Vaz-Velho, M. D/P/134<br />

Velge, P. B/P/47<br />

Verran, J. D/P/132<br />

Vidić, B. D/P/177<br />

Vignaud, M. C/P/82<br />

Virgilio, S. C/P/87<br />

Vishal Singh, P. C/O/30<br />

Vongkamjan, K. D/P/144, D/P/160<br />

Voronin, M. C/P/95<br />

Wadge, A. E/PL/13<br />

Waidmann, M. S. B/P/50, B/P/51, D/P/148<br />

Wal<strong>to</strong>n, J. D/P/193<br />

Wang, J. B/P/63<br />

Wang, W. A/O/13<br />

Warke, S. A/P/07, B/P/49<br />

Weiss, S. B/P/81<br />

Whitehead, K. D/P/132<br />

Whiting, R. C. A/P/16<br />

Wiedmann, M. A/PL/02, A/O/13, A/P/16,<br />

A/P/31, A/P/33, D/P/160, D/P/144<br />

Wilhelm, T. C/P/122<br />

Williams, D. B/O/14<br />

Winkelströter, L. K. D/P/148<br />

Wong, H. T. L. A/P/39<br />

Wren, B.W. A/P/05<br />

Xayarath, B. B/P/57<br />

Yurov, D. A/P/34<br />

Zahle Andersen, A. A/P/24<br />

Zaytseva, E.A. B/P/73<br />

Zeiman, E. A/O/10<br />

Zeppa, G. D/P/150<br />

Zhang, W. A/O/08<br />

Zuliani, V. D/P/154<br />

Zurbriggen, A. B/P/52<br />

Zweifel, C. D/P/131


Silver Sponsors


Bronze Sponsors<br />

Other Sponsors<br />

Scientific Sponsors


Universidade Católica Portuguesa – Escola Superior de Biotecnologia<br />

Por<strong>to</strong>, 2010

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!